Sie sind auf Seite 1von 198

http://www.handprint.com/ASTRO/ae1.

html

Astronomical Optics
Part 1: Basic Optics
The Geometry of Light
Refraction
Reflection
Wavelength & Frequency
Interference
Ray Tracing the Optics
Gaussian Concepts
Sign Conventions
Locating the Principal Planes
Image Attributes
Types of Lenses
Image Size & Location (Positive Lens)
Image Size & Location (Negative Lens)

Ray Tracing a Spherical Mirror


Lens Combinations
Thin Lens Formulas
Single Lens Optical Analysis
Multiple Lens Optical Analysis
Eyepiece Prescription Data
Optical Materials
Optical Coatings

This page introduces the optical principles necessary to understand the design and function of
telescopes and astronomical eyepieces. Subsequent pages discuss the telescope & eyepiece
combined, eyepiece optical aberrations, eyepiece designs and evaluating eyepieces.
Included at the end of each page is a list of Further Reading that identifies the sources used
here and background information available online.
The Geometry of Light

Light propagates in the form of oscillations in an electromagnetic field, which expand from the
light source as evenly spaced and concentric wavefronts whose energy is measured in the
quantum packets known as photons. Considered as photons, light travels in a straight line from
the light source unless or until it is absorbed, reflected or refracted by matter.
The radiation of light through space can be represented in two ways: (1) as wavefronts that
expand concentrically and radially from the light source (analysis by physical optics), or (2) as
straight lines originating in the light source and perpendicular to the wavefronts that indicate the
direction in which each part of the wavefront is moving (analysis by geometric optics). Basic
optics is entirely developed in terms of geometric optics.
In all astronomical applications, light sources are so distant that the concentric wavefronts are
effectively parallel planes over any practical telescope aperture, and the geometric rays
describing the direction of the wavefronts are also parallel. To illustrate: across the aperture of a

254mm (10") telescope, light rays from a single point on the moon, the closest astronomical
object, diverge from parallel by no more than 0.00000000066 millimeter, or 1/1,000,000 a
wavelength of light. Therefore, geometric optics can assume perfectly flat and parallel
weavefronts from a distant light source, and light rays that are straight, parallel and
perpendicular to the wavefronts.
Refraction
Refraction is the characteristic behavior of light that is incident on the smooth surface of a
transparent (transmitting) material.
Light has a uniform and universal speed in a vacuum of nearly 300,000 kilometers per second.
Light can also propagate through various transparent materials, such as air, water or glass, but
each material slows the speed of light by a specific value in some materials, to almost one
third its vacuum speed. The speed of light in a vacuum divided by the speed of light in a
refracting material is the refractive index (n) of the material: this is 1.00029 for air; 1.3333 for
water; and anywhere from 1.4 to 2.0 for optical glasses.
When light crosses the boundary between materials of different refractive index, the average
direction of the wavefront can be deflected in a new direction. The geometrical analysis of this
behavior is nicely summarized in a classic illustration made by Christiaan Huygens in 1678
(diagram, left). The physical wavefronts are indicated as parallel white bands, and the geometric
light rays as thin white lines.
As the wavefronts of light AB, traveling across distance BC, encounter a refracting boundary AC,
the speed of the wavefront is slowed so that it now travels a shorter distance AB' in the same
time. The refracted wavefronts are also turned in a different direction, which occurs because the
same wavefront encounters the boundary at different times (t1 to t5) across its width AB.
Because geometric rays are always (by definition) at right angles to the wavefronts they
describe, they create the right triangles ABC before refraction and AB'C after refraction, with
side AC in common. Inspection shows that the angle of incidence (1) is equal to the angle BAC,
whose sine is equal to BC/AC; and that the angle of refraction (2) is equal to the angle ACB',
whose sine is equal to AB'/AC. Since AC is a common denominator, the sines differ in the ratio
BC/AB'. This ratio is the proportional slowing in the refracted wave, which we have seen is
defined as the index of refraction, and therefore the sine ratio is equal to the inverse refraction
ratio n2/n1.
These relationships are summarized as Snell's Law or the Law of Refraction, illustrated in the
diagram by the yellow arrows and defined mathematically as:
sine(1)/sine(2) = n2/n1
or
sine(1)n1 = sine(2)n2
where n1 and n2 are the refractive indices of the two media that form the refracting boundary,
and 1 and 2 are the angle of incidence and angle of refraction. These angles are measured from
a line normal (perpendicular) to the boundary surface of the two media at the incidence point of
a light ray. Both light rays and the line normal must lie in a single plane, and the incident and
refracted rays will be on opposite sides of the line normal.
This analysis can be explaned in terms of the wavefront character of light, as the formation of
single wavefronts by the coalescing wavefronts expanding from each incidence point (curved arcs
along the line B'C). However Snell's law equally applies if the wavefront is summarized as a
single light ray parallel to the direction of the light and perpendicular to the wavefront (the
yellow arrows). The angles of incidence and refraction within a single plane remain the same.
Reflection
Reflection is the characteristic behavior of light that is incident on a smooth and opaque
(reflective) material. In this situation the Law of Refraction must be qualified in three ways:
For angles of incidence equal to 0, the ratio of the sines is zero and no refraction occurs: the
light is slowed but its direction is not altered.
If the angle of incidence is greater than a critical angle (CR), defined as:
CR = arcsin(n1/n2), n2 > n1

then the ray is reflected from the boundary rather than refracted through it. In a boundary with
air, the critical angle is about 49 for water (n = 1.33) and 46 to 30 for optical glasses (n = 1.4
to 2.0).
When reflection occurs, either from reflecting surfaces such as mirrors or glass surfaces
positioned to the incident light at angles greater than the critical angle, the angle of reflection
equals the angle of incidence but on the opposite side of and in the same plane as the line
normal:
2 = 1
For all angles of incidence between the critical angle and 0, significant reflection also occurs
when the difference between the refractive indices of the two media is greater than about 0.25.
This light energy is emitted by the surface at the reflected angle of incidence, again with the two
angles on opposite sides of the line normal.
Wavelength & Frequency
The distance between identical points on two
adjacent wavefronts of light is the wavelength () of
the light. The frequency () of light is the number of
wavefronts that pass a fixed point in one second, or
the cycles per second. The relationship between
frequency and wavelength is governed by the speed
of light in a vacuum, c:
c = = 299,792,458 ms1 = 3 x 108 m/s.
m = c / .
Light wavelengths are commonly measured in
angstroms (10-11 meter), nanometers (10-9 meter)
or micrometers (10-6 meter). Light that appears
green to the human eye has a wavelength of about
550 nanometers 0.55 micrometers or 0.00055
millimeters. Visible wavelengths range from about
750 nm ("orange red") to 380 nm ("blue violet"), or
0.00075 to 0.00038 millimeters. Longer wavelength
"red" light carries less energy, and therefore has a
lower frequency, than shorter wavelength "violet"
light.

Transparent materials do not refract all wavelengths of light at the same angle. Instead, higher
energy (higher frequency) "violet" light is refracted at a greater angle than lower energy "red"
light. This is the reason that prisms (or raindrops in the air) spread "white" light into the
characteristic light spectrum. Variation in the refractive index across different wavelengths is
called dispersion, discussed below under optical materials.
Interference
Although the optical behavior of light can be described in terms of geometric rays, the image
forming behavior of light requires the description of wavefronts in terms of physical optics. The
most important behavior is the interaction between light wavefronts that produces interference.
Interference can arise in two ways.
The first is diffraction or the concentric expansion of wavefronts from an occluding edge. This
was demonstrated by the slit experiment of the English naturalist Thomas Young which proved
the wave nature of light (diagram, left).
In this situation light composed of parallel wavefronts from a single light source encounters a
thin opaque barrier divided by two parallel, closely spaced and very narrow slits. The slits allow a
part of the wavefront to pass, but as it does so the wavefront expands concentrically from each
slit aperture.
The slits are positioned so that the oscillations of separate wavefronts must be either coincident

or opposed. Where they coincide a wavefront reinforcement occurs and a band of light will be
projected onto a screen placed behind the slits. Where they are opposed a wavefront cancellation
or destructive interference occurs the peak of one wavefront is cancelled by the trough of the
opposing wavefront and this produces a dark band on the projection screen.
The interference effects produced by physical obstructions are the origin of the diffraction
artifact produced by a star or "point" light source, and the bright and dark pattern of speckles
produced in the image of a star disrupted by atmospheric turbulence.
The second way in which interference can arise is through the reflection of the wavefront from
separate transparent media. This was first investigated by Isaac Newton as the interference
fringes that appear between two closely spaced pieces of glass. In the optimal situation, the
spacing between the first and second reflective surfaces is equal to 1/4 wavelength of the light.
Light is partially reflected from the first and second boundaries, and these separate wavefronts
are exactly 1/2 wavelength out of phase. The peak of one wavefront mixes with the trough of the
second, and the result is that the two wavefronts cancel each other and the light is destroyed.
This laminar interference is the physical cause of iridescent colors from laminated, pearlescent
materials such as abalone shells or bird feathers, or from a thin film of oil on the surface of
water. It is also the basic mechanism exploited in reflection reducing optical coatings on the
components of astronomical eyepieces.
Ray Tracing the Optics
Furnished with the geometrical description of light, the Law of Refraction and information about
the refraction and dispersion of optical materials, we can analyze the basic attributes of any
optical system. Optics can be divided into two levels of geometrical analysis:
First order analysis was developed by Carl Friedrich Gauss in his Dioptrische Untersuchungen
(1841), expanding on analyses by Isaac Newton, Johannes Kepler and Huygens. It deploys
Snell's law and a simplified trigonometric analysis to determine the focal length, magnification
and power of an optical system, which yields the location, size and orientation of the image it
creates.
Third order analysis was developed through the combined efforts of several 19th century
mathematicians and opticians to describe optical aberrations, or departures of focused light
from the optimal image location and size defined by the first order analysis.
First order analysis is developed from the simplified properties of paraxial light rays. These rays
are refracted by the surface of a lens very close to its vertex or intersection with the optical axis.
In that tiny area the refraction of light by the curved surface of a lens can be diagrammed as the
refraction of light by a plane perpendicular to the optical axis. This allows the arithmetic
calculation of image size and location without the use of trigonometric functions (sine or
tangent).
This paraxial approximation can usefully describe the optical behavior of moderately curved
lenses with a field angle of up to 20 if we accept errors of calculation of 1%, or to lenses of 6
with errors of 0.1%. Exact analysis is possible with trigonometric methods and finally by
calculating the physical properties of light wavefronts, but the paraxial approximation is
invaluable to describe the basic attributes of an optical system.
Gaussian Concepts
In the Gaussian analysis, the optical system is assumed to provide a perfect (distortion free and
precisely focused) image at the optical axis: analysis is only used to define the location, size and
orientation of this perfect image.
The analysis builds on the fact that the behavior of an optical system can be diagrammed in
relation to three pairs of cardinal points: the focal points, the principal points and the nodal
points. However, the nodal and principal points exactly coincide for lenses or mirrors surrounded
by air the standard situation in astronomical optics so only the focal and principal points are
needed to describe the system optical behavior.
The diagram (below) illustrates the key concepts and terminology in the first order analysis of a
schematic biconvex lens surrounded by air. These concepts can also apply to single or compound
lenses of any type, treated as a single optical unit or "black box".

A few basic properties of the optical system are assumed to apply. All optical components are
constructed as solids of rotation, which means their refracting surfaces are symmetrical around
an axis. The axes of rotation for all surfaces are identical with a single optical axis when light is
passed through the optical system. The intersection of a refracting surface with the optical axis is
the vertex of the surface.
Lens surfaces are assumed to be (and in nearly all real situations are) manufactured as sections
of a sphere, defined by a radius of curvature originating from a center of curvature located on
the optical axis. A two sided lens has two centers of curvature (denoted r1 and r2) and two radii
measured along the optical axis from the corresponding vertex. If one side of the lens is a flat
(plane) surface, the radius of curvature is zero.
Light rays arise from an object or object space (e.g., area on the celestial sphere) intersected by
the optical axis and conventionally positioned to the left of the lens. These rays pass through the
lens from left to right and terminate in an image or image plane located on the right of the lens;
the light receptor (the observer's eye, a CCD chip, photographic film) is therefore located at the
right of the lens oriented toward the left.
An object ray originates from some point on the object or object space being imaged by the
optical system; an image ray terminates at a corresponding point in the image of the object. All
image points are located on an image plane perpendicular to the optical axis and intersecting the
optical axis at the focal point. (Note that all real optical images are in fact focused onto a surface
that is more or less spherical or paraboloid, with its own radius of curvature; the image plane is a
paraxial simplification.)
Finally, all refracting optical systems are reversible: they can focus light passing through them
from left to right or from right to left. This creates a focal point on each side of the lens. The
object and image, and matching distances and points connected with them, are conjugate.
This is the basis for a number of related concepts, specific labels and symbols, illustrated in the
diagram (above):
Collimated rays are parallel to the optical axis and to each other. Oblique rays or abaxial rays
are at an angle to the optical axis (originate from an object point not on the optical axis).
If a collimated ray from a point on the object is extended through the lens, and the
corresponding oblique image ray is extended back from the conjugate image point, they will
intersect in a principal plane perpendicular to the optical axis and intersecting the optical axis at
a principal point. All rays from the object and refracted to the image will intersect in the same
plane. (Again, aberration free optical systems refract light as if from a spherical or paraboloid
surface; the principal plane is a paraxial simplification.)
In a thick or compound lens there are two principal points and corresponding principal planes.
The first principal plane, first principal point and first focal point are assigned to the surface
where light enters the lens; the second principal plane, second principal point and second focal
point are assigned to the surface where light exits the lens.
The effective focal length (EFL) is the distance from the second principal point to the second
focal point ('). The back focal length (BFL) is the distance from the back vertex of the lens to
the second focal point. The front focal length (FFL) is the distance from the front vertex of the
lens to the first focal point (). In most Gaussian equations the front and back focal lengths are
assumed to be equal, and multiples of the focal length can be used to specify the location of
images behind the lens and/or objects in front of the lens.
A chief ray or principal ray arises from any object point not on the optical axis, passes through
the first principal point, exits from the second principal point at the same angle to the optical
axis, and intersects the image plane at the conjugate image point.

An axial ray or marginal ray originates from the point where the optical axis intersects the
object or object space, enters the first principal plane at the outer edge or aperture radius of the
lens, then exits from the second principal plane to intersect the optical axis at the focal point.
An object ray intersects the first principal plane at a specific incidence aperture height (y) or
distance from the optical axis, and exits from the second principal plane at the refraction
aperture height (y').
The perpendicular distance from the optical axis to the most extreme object point is the object
height (h), and the distance from the optical axis to the most extreme (conjugate) image point is
the image height (h'). If the ray is collimated then the aperture height of the incident ray
corresponds to the object height of the originating point.
For elements notated by the same letter symbol before and after refraction, elements on the
image side, and rays or angles of rays after refraction by any surface, are denoted by an
apostrophe (e.g., and ', y and y', h and h').
In first order analysis all angles are measured in radians; for the very small slope angles
produced by paraxial rays, the tangent and sine of an angle are equal to the angle itself.
In astronomical optics most converging mirrors and a few wide angle eyepieces utilize aspheric
surfaces such as ellipsoids, paraboloids or hyperboloids (respectively solids of rotation generated
by an ellipse, parabola or hyperbola rotated around its major axis). In first order analysis these
are approximated as a spherical surface with a radius that produces the equivalent focal length.
Sign Conventions
In order for Gaussian equations to produce correct numerical values in all variations, the
algebraic sign of measured quantities used in the equations must follow arbitrary but specific
sign conventions:
As explained above, light is represented as radiating left to right; the object is always to the
left of the lens and the real image is always to the right.
A measurement reference point, usually the principal point of the optical system, is used as the
origin of a Cartesian coordinate system. Lengths measured along the optical axis from left to
right indicate the direction of incident light and are positive in sign; lengths measured right to left
are negative. Lengths measured upward from the optical axis are positive; lengths measured
downward are negative.
Angles measured at the optical axis are positive if the oblique ray must be rotated clockwise to
reach the optical axis, and negative otherwise. Angles of incidence and refraction are positive if
they must be rotated clockwise to reach the line normal of the refracting surface, and negative
otherwise. (Rotation is always through the smaller angle to the optical axis or line normal.)
Both focal lengths of a converging lens are positive, and both focal lengths of a diverging lens
are negative. The single focal length is positive for a concave (converging) mirror and negative
for a convex (diverging) mirror.
A radius of curvature is positive if it lies to the right of the surface it describes, negative if it lies
to the left.
Application of these conventions is illustrated in the optical calculations below.
Locating the Principal Planes
Before developing the Gaussian formulas for optical systems, it will be helpful to illustrate how
Snell's Law and basic trigonometry are applied to locate the second principal plane, second
principal point and effective focal length of a symmetrical biconvex lens (diagram, below).
In this example, the lens has a refractive index of nL = 1.6, surrounded by air with a refractive
index of nA = 1.0. We follow the path of a single collimated light ray (parallel to the optical axis
of the lens) that is incident on the front surface of the lens at an aperture height y from the
optical axis.
The lines normal to the front surface of the lens are produced as lines radiating from the second
center of curvature (r2, shown in the diagram on the far right). Since the lens is symmetrical, the
first center of curvature (r1) is of the same length and outside the diagram on the left.
The angle of the line normal to the optical axis at the entrance point i is the arcsine of y/r2.

Because we assume the lens produces a perfect focus, the aperture height y can be set to any
reasonable value for the analysis. In the example, I've chosen y = 10 and r2 = 30.9: the
incidence point then defines a sine of 0.309 to the line normal, giving an angle of entrance
incidence of 1 = 18. (Note the application of the sign conventions to the angle quantities.)
The ratio between the indices of refraction between the first and second materials (air and glass)
is 1.0/1.6 = 0.625, so the sine of the refracted ray is 0.3090.625 = 0.193, for an angle of
refraction 2 = 11 on the opposite side of the line normal, or 7 in relation to the optical axis.
A small amount of light is reflected from the lens surface at an angle 1 to the line normal (this
is, for example, why we can see reflections in storefront windows); the rest of the light ray
continues along the refracted path inside the lens.
The next step is to calculate the thickness of the lens along the refracted light path. This is a
function of the relative curvature of the front and back surfaces, the separation t between the
two vertices at the optical axis, and the refractive index n1 of the glass. That calculation is
explained below; for now, we identify the exit point i' by construction in order to focus on
finding the focal point and second principal plane.
When the light ray reaches the back surface of the lens at point i', the exit aperture height y' =
9.2 and the line normal drawn from r1 is 17 to the optical axis. Then 3 = 24 from the interior
ray to the line normal, for a sine of 0.407. The ratio of refraction indices is now inverted as
1.6/1.0 = 1.6, so the sine of the refracted angle is 0.4071.6 = 0.651, which is 4 = 41 to
the line normal outside the lens or 24 to the optical axis.
On this oblique path the image ray continues until it intersects the optical axis at the effective
focal point ' with an image ray slope of u' = 24. (To avoid misunderstanding, note that only
a collimated ray parallel to the optical axis will intersect the optical axis at the effective focal
point.)
Diagrammatically, the original light ray can be continued from point i as a straight line parallel to
the optical axis, and the image ray can be extended backward from point i', until the two rays
intersect at point I at aperture height y. The angle of the image ray slope is then defined as
u' = y/' radians. Since y is given and u' has been calculated, the focal length is calculated as
' = y/u'.
A plane through the point I and perpendicular to the optical axis is the principal plane of the lens,
and the intersection of this plane with the optical axis is the principal point of the lens. The
effective focal length (') is simply the distance along the optical axis between the second
principal point and the focal point.
Image Attributes
At this point it is useful to introduce the terminology for the four attributes of an optical image.
(1) An image is real if the image is formed at a location where light rays have actually
converged: the sun concentrated by a magnifying glass or light focused by a photographic lens.

Real images can affect photosensitive media and can be projected onto a surface.
An image is virtual if the image appears at a location where no light is focused: for example, the
image in a mirror. Virtual images occur when light rays form a focus when extended backward;
they are formed by negative lenses or by placing an object inside the focal length of a positive
lens. A virtual image has no location in physical space and cannot affect photosensitive media.

(2) An image is erect if it is oriented vertically in the same way as the object, or inverted if the
image is reversed top to bottom.
(3) An image is normal if it is oriented left to right in the same way as the object appears, or
reverted if the image is reversed left to right. (Images rotated 180 in relation to the object are
both inverted and reverted.)
(4) Finally, an image can be enlarged, actual size or reduced in comparison to the physical size of
the object or to the angular width of the object as it appears to the unaided eye.
As the diagram shows, nearly all astronomical telescopes rotate the object image; none actually
invert the image, despite the common use of inverting telescope to describe them. Note also that
an inverting eyepiece actually produces an erect, normal image: the image is rotated by the
objective, not the eyepiece. The exception is the Gregorian telescope, because it reflects light
from a secondary placed beyond the focal point of the primary mirror. This rotates the image a
second time.
Types of Lenses
The possible combinations of spherical and plane surfaces that can be used to make a lens fall
into six generic types of positive or negative lenses, shown in the diagram (below) with two
indications of their relative refracting power: the effective focal length and a schematic ray
tracing. These illustrations show lenses of a typical optical glass (nL = 1.6) in air; nominal
radius for the strong curvature (in the meniscus lenses) is r = 50mm and for the weak curvature
(all other lenses) is r = 100mm in an aperture diameter Do = 70mm.

The biconvex, positive meniscus and plano convex designs are positive lenses used to
concentrate parallel rays of light to a single focus located behind the lens or to reduce the focal
length of an optical system. The biconcave, plano concave and negative meniscus designs are
negative lenses used to expand parallel rays of light away from a negative focal point (located in
front of the lens), typically to extend the focal length and increase the relative aperture of an
optical system.

The effective principal point is shown by the red dot. Note that even when symmetrical, "thick"
lenses have two principal points for the refractive effect in opposite directions. In general, the
principal points in a biconvex lens are both internal and spaced about 1/3 the distance from the
front to back vertices. In a lens with one plane surface, the first principal point is the front vertex
of the lens. In a meniscus lens, one or both principal points can be external to the lens.
The biconvex and biconcave lenses have roughly twice the refractive power of the plane and
meniscus forms. The meniscus lenses are most sensitive to changes in the lens thickness, while
the plano convex and plano concave lenses are unaffected by variations in thickness. In
asymmetrical positive lenses, the power of the lens is greater when the surface with the shorter
radius (or the curved surface in plano lenses) is oriented toward the object. In asymmetrical
negative lenses, the reverse holds: the power is greater when the greater power or curved
surface is oriented away from the object.
Used as single lenses, positive lenses always produce real, inverted, reverted images, while
negative lenses produce virtual, erect, normal images.
Positive meninscus lenses are commonly used in eyeglasses to correct for farsightedness or
difficulty focusing on objects at short distances from the eye (hyperopia). This occurs because
the lens to retina focal distance of the eye is too short, or the cornea and lens of the eye lack
optical power, or the lens has hardened with age and cannot adjust to focus on near objects
(presbyopia). Negative meninscus lenses are used to correct nearsightedness or difficulty
focusing on objects at long distances (myopia), which occurs because the cornea and lens are too
strongly curved, or the lens to retina distance in the eye is too large.
Image Size & Location (Positive Lens)
Now that the underlying logic of first order analysis has been illustrated and the concepts
defined, we can turn to the basic formulas that result first for positive or converging lenses,
then for diverging or negative lenses.
In the Gaussian model, the optical effect of a lens can be analyzed through the use of three
analysis rays. The diagram below shows this analysis applied with two principal planes, which is
done by disregarding the space between them. This is equivalent to replacing both planes by a
single principal plane midway between them, in what is called the thin lens model of the optics.
If it is acceptable to assume that the optical effect of the lens thickness (the distance between
the front and back incidence points of a light ray) is inconsequential to the slope of the exiting
image ray, then the lens can be modeled by a single principal plane located at the center of the
lens. This directly yields the effective focal length (measured from the single centered principal
plane) as:
1/' = (nL1)(c1c2).
where c = 1/r. For a symmetrical biconvex lens where r1 = 10 cm, r2 = 10 cm and nL = 1.6:
1/' = (11.6)((1/10)(1/10)) = 1/(0.62/10) = 1/0.12, or ' = 8.33 cm.
(Note application of the sign conventions.)
The "thin lens" analysis originated in 18th century optics in which the main applications were low
power, long focal length spyglass lenses, simple eyepieces and very thin eyeglass lenses. It is
applicable to any lens where the focal length is much larger than its maximum thickness. That
criterion is ambiguous and only suggests how well a physical lens might be described by the
idealized thin lens model: in particular, it implies that wide angle or short focal length (strongly
curved) lenses cannot be analyzed in this way.
Given that we have already located the principal points and focal points from the formula above,
we are interested to find the location, size and orientation of the image formed by a specific
object at a distance s from the lens.
From a single object point off the optical axis at object height h, construct:
A A collimated light ray (parallel to the optical axis) from the object point to the first principal
plane of the lens, and exiting from an equal height in the second principal plane as an oblique ray
through the second focal point '.
B An oblique ray from the object point through the first principal point p1 in the first principal
plane, and exiting from the second principal point p2 at the same angle to the optical axis.

C An oblique ray from the object point through the first focal point to the first principal plane,
and from an equal height in the second principal plane as a collimated light ray (parallel to the
optical axis).
Note that only two rays, A and C, are necessary to define the image; ray B is shown as a dashed
line because it is optional. Rays A and C only require the location of the two focal points, given
by the formula above. For an object at an infinite optical distance the distance at which all
rays from the object are parallel to the optical axis the use of ray C becomes invalid, because
to intersect it must be identical to the optical axis. In that case the image is identical with the
focal point intersected by ray A and the size of the image is determined only by the focal
distance multiplied by the field angle (in radians) between the optical axis and the object point
emitting ray B.

For an object at finite optical distance, all three rays will intersect at a single point that is not on
the optical axis. Given the first and second focal distances and ' and object height h, this
construction defines the object distance s, the image focal distance s', the to object distance x,
the ' to image distance x', and the image height h'.
The image location is the distance from the second principal point to a line through the ray
intersection that is perpendicular to the optical axis; the image size is determined by the distance
of the intersection from the optical axis, and the image orientation is shown as the interesection
lying above or below the optical axis (indicated by a positive or negative sign in the calculation of
the image height or the lens magnification).
Note the explicit symmetry in the procedure: an object height h at distance s produces an
image height h' at distance s', but an object height h' at distance s' would produce an image
height h at distance s. This symmetry is characteristic of conjugate points (joined in a
reciprocal relationship) in an inverting optical system.
Reasoning from the various congruent triangles created by the three rays, and given the object
distance s and focal length , one can derive the thin lens formula:
1/s' = 1/s + 1/'
and the other parameters:
[ to object distance] x = s+
[' to image distance] x' = ('2)/x
[object distance] s = s''/('s')
[focal distance] s' = s'/(s+')
[object size] h = (h'x)/'
[image size] h' = (hx')/'
and several equations for the image magnification (m), which in the Gaussian analysis refers to
the ratio of the linear size of the image over the physical size of the object:
m = h'/h = '/x = x'/' = '/(s+')
The curvature (c) of the lens is the reciprocal of the radius of curvature; the power () of the
lens is the reciprocal of the effective focal length:
c = 1/r
= 1/'
Refractive power can be expressed in a standardized measure, the diopter, which is the
reciprocal of the focal length in meters (1/m). The shorter the focal length, the more the
objective or eyepiece has refracted (or reflected) light, and the higher the diopter. Thus the
diopter for an objective with o = 500mm is 1/0.5m or 2; the diopter of an eyepiece with focal

length e = 20mm is 1/0.02m or 50.


As an illustration and check on your practice calculations, the relationships among these different
measures are illustrated in the table below for a constant object size of 2.5 cm and a
symmetrical biconvex lens with an effective focal length of 25 cm.
object
distance (cm)
s

s/

to object
distance (cm)
x

object
size (cm)
h

focal
' to image
distance (cm) distance (cm)
s'
x'

image
size (cm)
h'

magnification
m

0.02

0.50

24.50

2.5

0.51

25.51

2.55

1.02

0.10

2.50

22.50

2.5

2.78

27.75

2.78

1.11

0.20

5.00

20.00

2.5

6.25

31.25

3.13

1.25

0.50

12.50

12.50

2.5

' = 25.00

50

5.00

2.00

0.67

16.75

8.25

2.5

50.76

75.76

7.58

3.03

0.80

20.00

5.00

2.5

100

125

12.5

5.0

0.90

22.50

2.50

2.5

225

250

25

10

0.99

24.75

0.25

2.5

2475

2500

250

100

1.00

' = 25.00

2.5

1.01

25.25

0.25

2.5

2525

2500

250

100

1.10

27.50

2.50

2.5

275

250

25

10

1.25

31.25

6.25

2.5

125

100

10

4.0

1.5

37.50

12.5

2.5

75

50

5.0

2.0

50.0

25

2.5

50

25

2.5

1.0

125

100

2.5

31.25

6.25

0.625

0.25

10

250

225

2.5

27.78

2.78

0.278

0.111

50

1250

1225

2.5

25.51

0.51

0.051

0.020

500

12500

12475

2.5

25.05

0.05

0.005

0.002

5000

125000

124975

2.5

25.01

0.005

0.0005

0.0002

50000

1250000

1249975

2.5

' = 25.00

0.0005

0.00005

0.00002

Again, note carefully the sign conventions necessary for consistent calculations:
for a converging lens both focal lengths are positive, for a diverging lens both focal lengths are
negative;
s is measured from the first principal plane and is always negative;
x is measured from the first focal point and is negative if the object is in front of the first
focal point or positive if the object is behind it;
the focal distance s' is negative if measured from the first principal plane and positive if
measured from the second principal plane;
x' is measured from the second focal point ' and is positive for a real image (behind the lens)
and negative for a virtual image (in front of the lens);
when x' is negative, the image location measured from the first principal plane is equal to
x'+'.
If only one principal plane is used in the ray tracing (the "thin lens" model), then all distances are
measured from it.
This analysis method indicates the changes in the image location, size and orientation produced
by objects at different distances from the lens (diagram, above):
(a) Objects that are far distant enough to make the principal rays A, B and C effectively parallel
are said to be "at infinity", and will form an inverted image at the focal point '. If we take the
minimum discernable image divergence from the focal point as equal to a wavelength of light
(~0.0005 mm), then an object distance of roughly 50,000' (e.g., 12.5 kilometers for an
'=25cm lens) is "at infinity" for calculation purposes, with a magnification of /50,000.
When the object is between 50,000 to 2 in front of the lens, the image will appear inverted
between 1 and 2 behind lens with a magnification between '/50,000 and 1.

(b) When the object is located exactly 2 in front of the lens, the image will appear inverted 2
behind the lens (s' = s) and actual size (h' = h, m = 1).
(c) If the object is located between 1 and 2 in front of the lens, the image will appear
inverted at more than 2 behind the lens, and larger than actual size (m > 1).
(d) If the object is located exactly at 1 no image can form (the ray C through the front focal
point never reaches the principal plane).
(e) If the object is located at a distance less than 1 then the analysis rays must be extended
backward from their intersection points with the principal plane. They intersect in front of the
first principal plane to form an erect (h'>0) and enlarged (m > 1.0) virtual image. Near the limit,
where s approaches zero, the image forms on the object itself and is actual size.
Note that it is not possible for an object at any distance to form an image between the lens and
the second focal point (at a distance less than the effective focal length).
Image Size & Location (Negative Lens)
The diagram below shows the "thin lens" analysis for a negative (diverging) lens, again with two
principal planes; the analysis is the same if the two planes are replaced by a single plane.
The major difference from the positive lens analysis is that the effective focal length is measured
in front of the lens, and both focal points have negative distances.

From a single object point off the optical axis at object height h, construct:
A A collimated light ray (parallel to the optical axis) from the object point to the first principal
plane of the lens, and exiting from an equal height in the second principal plane as an oblique ray
(yellow line) that when extended backward as a virtual ray (magenta line) passes through the
first (effective) focal point '.
B An oblique ray from the object point through the first principal point p1 in the first principal
plane, and exiting from the second principal point p2 at the same angle to the optical axis.
C An oblique ray from the object point to the first principal plane that when extended forward
from the second principal plane intersects the second focal point , but is extended forward from
the second principal plane as a collimated light ray (parallel to the optical axis) and is extended
backward as a collimated virtual ray.
Given the first and second focal distances ' and and object height h, this construction defines
the object distance s, the image focal distance s', the to object distance x, the ' to image
distance x', and the image height h' as shown in the diagram (above).
A real image cannot be produced by the diverging refracted light rays (yellow lines): instead a
virtual image is formed at the intersection of the virtual rays (magenta lines) extended backward
from the emergent light rays. For an object at finite optical distance, all three virtual rays will
intersect at a single point that is not on the optical axis. For an object at infinite optical distance
the image is again located by means of only rays B and C from an off axis object point.
The focal points both have negative distances from their respective principal planes, which
means (according to the sign conventions and the thin lens formulas) that the focal distance
s' is always negative and the ' to image distance x', the image size h' and the magnification m
are always positive. (A virtual image cannot form outside the effective focal length '.)
This analysis shows that a negative lens always produces an erect but virtual image when viewed
from the side opposite the object. For an object at infinity, analysis ray B shows that the virtual
image forms at the first focal point ' with a height equal h' = radians()'. As the object is
brought closer to the lens, the virtual image becomes larger and also approaches the lens; if the
object is located at the first focal point then the virtual image is located at s' = /2 and at half

size (m = 0.5).

Ray Tracing a Spherical Mirror


In a mirror there is only one optical surface, so analysis is simplified. The major variations are for
objects at finite or optically infinite distances, for concave or convex mirrors, and for real and
virtual images.
The diagram (below) shows the analysis for a concave mirror with the object at a finite distance,
and for a concave and convex mirror at infinite optical distance.
The sign conventions for a concave mirror are that the single focal point and image focal distance
are positive, because they are measured in the direction that the light is traveling; the object
distance is measured opposite the direction of the light and is therefore negative. For a convex
mirror the focal point and image focal distance are also negative. The object and image heights
and image angles are negative below the optical axis.

The image size and location are defined as with lenses, by means of three analysis rays from a
single off axis object point:
A A ray from the object point, parallel to the optical axis and reflected by the mirror.
B A ray from the object point through the mirror focal point and reflected by the mirror.
C A ray from the object point through the mirror center of curvature and reflected by the
mirror; this ray is reflected back through the center of curvature.
D A fourth ray, from the object point to the mirror vertex, can be used instead of ray C for
aspheric (ellipsoid, paraboloid, hyperboloid) mirrors that do not have a center of curvature.
The first diagram (above) shows the ray tracing for a concave (positive, converging) mirror with
an object point of height h at finite distance o from the mirror vertex. The tracing shows that all
three rays converge at a real image in front of the mirror, shifted from the focal point toward
the object at a focal distance i with a reduced image size h'. A line through this point and
perpendicular to the optical axis will indicate the location of the real image.
In contrast, an object placed inside the focal point of a concave mirror will produce an enlarged,
erect and virtual image behind the mirror. This is the standard design of a magnifying vanity
mirror, where the focal point is farther from the mirror than the typical viewing distance.
The middle diagram (above) shows the ray tracing for a convex (negative, diverging) mirror,
again with an object point at height h and at finite distance o from the mirror vertex. The tracing
shows that rays B and C are reflected before they pass through their target points behind the
mirror, so all three reflected rays are extended in the opposite direction as virtual rays behind
the mirror, where they converge at the location of a reduced, erect virtual image at focal
distance i which again is shifted from the focal point toward the object. Since the object cannot
be placed between the mirror and the focal point, no real image is formed. (The exception is
when the convex mirror reflects rays that are convergent, for example when used as the
secondary mirror in a two mirror telescope.)
The relationships between object distance, image distance and focal length are:
1/ = 1/o+1/i
1/i = 1/o1/.
The center of curvature and focal point are related as
= r/2
which means ray A scales the image size to the object size as
h' = hi/o
with magnification
M = i/o = h'/h.
The last diagram (above) shows the ray tracing for a concave mirror with an object at infinite
distance the usual astronomical situation. In this case all rays from the object are parallel, so
rays B and D are sufficient to define the image location and size. For off axis objects, both rays
define the angular width or field height measured from the optical axis or center of the image
field. Any collimated ray A will identify the focal point, which for a spherical mirror is found as
= r/2.
The image forms at the focal point , and the field height h' of any part of the image is equal to
the focal length times the incidence angle (in radians) from the object height to the mirror
vertex or through the focal point. This angle is opposite in sign to the original angle because it
is reflected on the opposite side of the optical axis:
h' = radians()
The average diameter of the Moon is approximately 30 arcminutes; its disk will be 21.8 mm wide
at the focal plane of a 250 mm /10 mirror.
The diagram also shows that as the aperture height y increases, the location of the mirror
surface shifts by a distance e from the vertex location due to the surface curvature. The surface
of rotation that focuses all rays parallel to the optical axis at the focal point is determined by a
parabolic function, so this shift is:
y2 = 4e and e = y2/4.
For a 250mm (10 inch) mirror figured to /4 ( = 1000mm), the depth of curvature required
from mirror edge to center (y = 125mm) is 3.9mm (0.15").

Single mirrors, like lenses, produce a curved focal surface whose radius depends on the focal
length, roughly as:
=
Lens Combinations
Astronomical eyepieces are designed as two or more lenses or components mounted along a
common optical axis. Their design and evaluation requires optical formulas that extend and
generalize the analytical framework described above for a single lens. This section first outlines
the traditional thin lens formulas, then more general analytical formulas.
Thin Lens Formulas
A number of optical formulas were derived in the 18th century to characterize the behavior of
"thin lens" combinations calculated in relation to a single principal plane for each optical element.
We start with this framework for its simplicity and to illustrate basic relationships between two
lenses and their spacing as an optical system.
The design procedure is first to select two lenses of the appropriate aperture and power (1 =
1/1 and 2 = 1/2). The lens spacing (the distance between their principal planes) is
determined by trial and error or by analytical formula, for example to minimize chromatic and
spherical aberration and optimize eye relief.
Once the interlens spacing is fixed, the front focal length (FFL) is calculated to determine the
location of the eyepiece focal plane and field stop. From this point the effective focal length (EFL)
and location of the eyepiece principal plane are determined. The back focal length (BFL) is
calculated to determine the location of the exit pupil and the approximate amount of eye relief.
Finally, the radius of the eyepiece apparent field of view is determined by the angle , drawn
from the center of the exit pupil to the edge field stop aperture projected onto the eyepiece
principal plane.
A worked example for a simple two element eyepiece, using the two second principal planes for
two components with focal lengths of 115mm (for the front or field lens) and 70mm (for the back
or eye lens) at a separation of 50mm, is shown in the diagram (below). This illustrates the
application of the formulas for front and back focal length (FFL and BFL) and effective focal
length (EFL). If the "thin lens" approximation is applied, the same formulas can be used with a
single principal plane through the center of each lens.

Note the following:


Assuming the thin lens model is valid, then the focal length of a single spherical lens can be
calculated from its index of refraction (nL) and the front and back curvatures (the lensmaker's
equation) as:
1/ = (nL 1)[1/r1 1/r2]
and the effective focal length of a compound lens (where d = 0) can be found as:

EFL = (12)/(1+2).
Recall that the focal length of a negative (diverging) element is negative.
The effective focal length determines the principal plane of the lens combination, and is
measured forward from the field stop. The EFL will typically be shorter than the focal length of
either of the lenses separately in positive (converging) lens combinations.
The apparent field of view (the angular diameter of the field stop viewed from the exit pupil) is
derived from the radius of the field stop (do) as double the exit angle of a marginal ray:
AFOV = 2*arctan(do/2*EFL) = 2*
To maximize eye relief, the lens with the higher power is usually used as the eye lens. (To
verify this, note that FFL is shorter than BFL.)
A point discovered by Huygens, and implemented in the eyepiece design that bears his name,
is that two lenses of the same refractive index but different powers produce the most achromatic
image when
d = (1 + 2)/2.
The Kellner modification of the Huygenian design handles chromatic aberration by using an
achromatic doublet for the eye lens.
The graph (below) shows the optical effect when the focal length of either the front or back lens,
or the distance between them, is reduced by 50mm while holding the other elements constant.

The effect on eye relief (red line) is the same as the focal length of either the field lens or eye
lens is made shorter. Higher power lenses yield less eye relief, but not as a proportion of the
focal length. Thus, the field lens focal length is reduced by 43% (from 115mm to 65mm), and
the eye lens focal length by 71% (from 70mm to 20mm), but each change reduces eye relief by
the same amount, from 34mm to 12mm. In contrast, reducing the lens spacing from 50mm to 0
increases eye relief, but only by 10mm.
Reducing the focal length of the field lens has a relatively small effect on the system effective
focal length (green line) and the system apparent field of view (blue line). This is in contrast to
the effect of a higher power eye lens, which sharply reduces both the focal length and apparent
field of the system, an effect that is augmented by a closer lens spacing.
The direction of these changes is the same in both lenses, and differs from the effect of the lens
spacing only in the change in eye relief, so the eyepiece focal length decreases and the apparent
field of view increases, and by a much greater proportion than the eye relief is reduced, when
both lenses are made to a shorter focal length and are placed closer together.
The diagram (below) shows the analysis for a "thick" lens eyepiece, the 28 mm Edmund RKE.
Again, the marginal "pencils" or beams of light are traced through the system, but now the
calculation of focal lengths is based on the pair of principal planes for each component; the
intervals between the principal planes are ignored.

These examples reveal the underlying design principle of astronomical eyepieces:


The function of the field lens is primarily to "stage" or prepare the image by partially correcting
the divergence in the light rays (caused by the relative aperture or ratio of the objective) after
the rays have passed through the focal plane of the eyepiece, refracting the rays from a
divergent to an approximately collimated beam. Additionally in the RKE (and all "wide field"
eyepieces), the field lens increases the radius of the beam so that a longer eye relief will deliver
the same apparent field.
The function of the eye lens is to bring these corrected rays into a much shorter focus at a
much steeper angle to the optical axis, which produces a wider apparent field of view and greater
magnification. At the same time, the eye lens eliminates any remaining divergence in the image
rays so that they exit the eyepiece as parallel bundles, termed afocal because the image is not
formed at a single focal point but remains in focus when projected to any distance.
The significance of the exit pupil is that it is the point where the beam compression of the system
is at maximum and has the smallest diameter (obvious in the diagram). This highly compressed
beam can most easily accommodate the small entrance pupil aperture of the observer's eye.
Both the field lens power and the lens spacing can be manipulated to compensate for the very
short eye relief that a high power eye lens will produce, to make alignment of eye pupil and exit
pupil comfortable for the observer.
The main constraint remaining is that a high power eye lens or a very wide field design (when
> ~20) produces significant optical aberrations, an optical design problem that was gradually
overcome by advances in optical theory and glass manufacture in the 19th and 20th centuries.
Single Lens Optical Analysis
Optical systems can be analyzed using the paraxial ideal by modifying the analysis based on
distances in relation to a single principal plane to one based on angles and field heights in
relation to multiple principal planes.
Again, the sign conventions are positive for measurements in the direction of light (left to right)
and for objects above the optical axis; negative against the light and below the optical axis.
Angles are always expressed in radians, which is approximately the tangent for angles less than
20.
First, let's revisit the procedure outlined above for locating the focal point and second principal
plane of a single converging lens. This can be generalized by analyzing an axial rather than
collimated incident ray (diagram, below).

The refracting power of lens in air results from the radius of curvature of the two faces of the
lens, the thickness of the lens, and the index of refraction of the glass. Therefore measurement
of the two surface curvatures (c1 = 1/r2 for the front surface, c2 = 1/r1 for the back surface) and
the distance between the two lens vertices (t) is sufficient to identify the second principal plane
analytically, using any feasible values for the distance s of an axial point on the optical axis and
for the aperture height y of an axial ray from that point to the lens (diagram, above).
For a single lens (where n0 = n2 = the index of refraction for air), the optical angles u1, u'1 = u2
and u'2 are measured in radians but now in relation to the optical axis or lines parallel to it
rather than to the lines normal defined by the lens surfaces.
The step by step calculations proceed as follows:
Given:
c1, c2 front, back curvatures of lens (=1/r1, 1/r2; r2 and c2 are negative by sign convention)
t vertex to vertex thickness of lens (positive, by sign convention)
n0, n2 index of refraction for air ( = 1.0)
n1 index of refraction for lens material
s distance of axial point (negative, by sign convention)
y aperture height of incident axial ray (positive when above the optical axis, by sign convention)
Calculate:
[1. entry incidence slope] u1 = y/s
[2. entry refraction slope] u'1 = [y(n1n0)c1+(n0u1)]/n1
[3. exit (image ray) aperture height] y' = y+[t(n1u'1)/n1]
[4. exit (image ray) slope, given u2 = u'1] u'2 = [y'(n2n1)c2+(n1u2)]/n2
[5. back focal length] BFL = y'/u'2
[6. effective focal length] EFL = y/u'2
[7. power of the lens] = 1/EFL.
This procedure can be repeated for a cemented compound lens consisting of two or three
elements. Simply make the aperture height and incidence slope of the entry ray into the second
(or subsequent) lenses equal to the exit ray aperture height and slope from the previous lens,
and calculate steps 2-4 with the radii of curvature of the second lens. (If the two lenses are
separated by air, then a different calculation is necessary, as described below.)
For example, in step 2 n1 replaces n0, n2 (the index of refraction for the following lens) replaces
n1, c2 is the curvature and y' replaces y, etc. The exit calculations 3-4 will use the final curvature
and the index of refraction for air. Note again that the angles (in radians) are assumed equal to
their tangents or sines, which is the essential condition for the paraxial approximation.
Once all calculations are completed, first the focal point is located by measuring the back focal
length from the back vertex as before. Then the principal plane of the compound lens is located
by measuring the effective focal length forward from the focal point. The power of the compound

lens is determined as before from the effective focal length.


To identify the opposite focal point, the two radii of curvature are reversed in the formulas and
the sign conventions will reverse the signs of the radii of curvature and ray angles.
Multiple Lens Optical Analysis
Next we generalize the "thin lens" formulas for two air spaced lenses or multiple element
components treated as single lenses. This procedure departs from the trigonometric analysis
given above in that the thickness of the optical components [single or compound lenses] is
disregarded, and replaced by a pair of principal planes. These are specifically planes of unit
magnification, which means, given an object placed at the first focal point, that:
Front and back focal lengths are equal: = '.
All rays entering the first principal plane at aperture height y will exit the second principal plane
at aperture height y' = y. (It is always assumed that the object and image rays lie in the same
longitudinal or lengthwise plane, called the meridional plane, which also includes the optical
axis.)
An axial object ray angle u will produce an image ray angle u', and an oblique object ray
angle u through the first principal point will produce an oblique image ray angle u' exiting
from the second principal point.
The image magnification will be 1.0, or h = h'.
The ray tracing requires only two analysis rays in a common meridional plane: (1) an axial
(marginal) ray from the object at distance s from the first principal plane of the field lens of the
system that passes through the principal plane at aperture height y (the radius of the system
aperture), and (2) an oblique ray from object height h that passes through the first principal
point of the field lens. Note that h is not necessarily equal to y. Then the object ray relationships
will be (diagram, below):
u = y/s and up = h/s (note that s and up are negative according to the sign conventions)
and the corresponding image ray relationships for the system will be:
u' = y'/s' and u'p = (h'y'p)/s (note that s', u' and u'p are negative by the sign
conventions)
As the diagram makes clear, the image angles have been transformed by the optical system such
that the simple equalities y' = y and u' = u no longer apply. So the generalized procedure must
get us from the object to the image values, using only the refractive indices n1 and n2 of the two
optical components, the spacing d between the facing principal planes of the two components,
and (for finite object distances) the total distance T between object and image.

First, we can get the generalized form of the "thin lens" equations of the previous section, which
will then apply to both thin or thick components. Given the unit magnification of the principal
planes for a single component yields the basic relationships:

u = y/s and u' = y/s'


Substituting these into the thin lens formula (above) by replacing object and focal distances by
ray angles (e.g., 1/s = u/y) and reciprocal focal length with power (1/ = ) yields as the
refraction of the first component:
u'1 = u1 y11
the transfer equations into the second component:
y2 = y1 +du'1 and u2 = u'1
the exit ray from the second component:
u'2 = u2 y22
and the effective focal length of the total system (components 1 and 2 combined at distance d):
EFL = u'2 /y1
Alternately, assume that 1, 2 and d are known or fixed by design, and we want to find the
focal length or power of the system:
sys = 1 + 2 d12
which is identical to the reciprocal of the focal length formula given above in the "thin lens" case:
sys = 1/1 + 1/2 d/12
If the effective focal distance, back focal distance and component spacing are known, the
separate effective focal distances of the components can be found as:
1 = (dEFL)/(EFL BFL) and 2 = (dBFL)/(EFL BFL d)
These equations permit fast layout and approximate analysis of optical systems without the
incremental calculation of ray traces through all refracting surfaces.
Finally, all focal optical systems conform to four systematic proportionalities that equivalently
define a quantity known as the optical invariant or Lagrange invariant (L). Given an axial ray
with a height y at the first principal plane and y' at the back principal plane, and an oblique ray
with height yp at the first principal plane and y'p at the back principal plane (diagram, above),
then the angles (in radians) of these rays to the optical axis will be u, u', up and u'p respectively.
In that case:
L = hn1u = h'nku'.
This quantity states the relationships between refractive power, object distance and image focal
distance purely in terms of angles and relative heights. It can be reduced to an alternative
statement of system magnification:
M = h'/h = n1u/nku'.
The invariant applies equally to single or compound lenses or multielement optical systems,
which are treated as a "black box" and analytically bracketed by the entrance pupil and back
principal plane.
In analyses related to optical aberrations, the first principal plane is usually coincident with the
entrance pupil of the system, which is the opening in the system that determines its aperture.
This is normally the mounting for the objective lens in a refracting telescope, the
circumference of the first mirror in a reflecting telescope, or the mounting of the corrector lens
in a catadioptric telescope.)
Eyepiece Prescription Data
In most modern eyepieces, the thin lens formulas are insufficient to design an eyepiece or guide
its manufacture. The total information required to specify an eyepiece optical design is called the
prescription data, which includes the radius of curvature for all lens surfaces, the distance
between lens surfaces and centers of curvature as measured along the optical axis, and the
specific glasses that provide the refractive indices and Abbe coefficients assumed to calculate the
light paths. A full example is shown below for a modern version of the symmetrical Plssl
design.

The prescription data are given in tabular form (diagram, lower left): measurements begin at the
objective focal plane and end at the eyepiece exit pupil, assumed to be coincident with the eye
pupil. Note that measurements are relative the radius r1 is measured from d1, distance d2 is
measured from d1, r2 is measured from d2, and so on. Negative numbers indicate measurements
toward the light source (usually, to the left), plane surfaces have a radius of 0, and air spaces
are shown as distances with no glass indicated. (This single radius format applies to spherical
lenses only; aspheric lenses require a polynomial prescription not shown here.)

Given the prescription data, available in eyepiece patent documents or optics references, and an
assigned diameter for the exit pupil (5 mm in the diagram) and object field diameter (equivalent
to the interior diameter of the eyepiece field stop), ray tracing software can calculate the path of
light from the objective focal plane through the eyepiece for any field height on the object focal
plane. These rays are often shown in different colors for field heights equal to 0%, 70% and
100% of the object field radius; the colors do not denote spectral frequencies but help to
interpret the diagram visually. Conventionally, three rays are calculated as radiating from each
field point so that they exactly fill the exit pupil.
Ray tracing allows calculation of the eyepiece principal plane, principal point, effective focal
length, back focal length, eye relief, the angle of the apparent field of view, and the Petzval
radius, which describes the field curvature in the focal plane of the eyepiece. (Note that a
negative Petzval radius indicates a positive field curvature the center of the focal surface is
closer to the field lens than the perimeter of the focal surface; a large Petzval radius describes a
relatively small field curvature.)
A single spectral frequency is necessary to compute these standard optical parameters; this is
usually = 550 nm or "yellow green" [called green] light. Analysis of eyepiece chromatic
aberrations requires ray tracing at different wavelengths, typically 475 nm ("blue"), 512 nm
("green"), 550 nm ("yellow green"), 587 nm ("orange yellow") and 625 nm ("red orange").
Optical Materials
Glass is the generic material used to refract light in astronomical instruments. Glasses are

amorphous (not crystalline) mixtures of fused silica (silicon oxide, SiO2) and oxides of several
metals including sodium, calcium, magnesium and aluminum added to improve the
hardness, durability and water resistance of the glass. It is relatively stable under normal
temperature changes, readily fabricated and economical.

Glasses, plastics and other transparent materials bend all wavelengths of light through refraction,
but bend low energy "red" light less strongly than high energy "violet" light, which spreads out
the refractive indices for different spectral hues in an optical effect called dispersion. Optical
materials are therefore identified according to those two attributes: (1) the index of refraction
and (2) the Abbe V number or measure of dispersion. In glass catalogs the two attributes are
reported as a six digit code the first three digits are the index of refraction (after the decimal
point), followed by the three digit V number (omitting the decimal point).
Across the entire electromagnetic spectrum, most glasses and plastics have periodic absorption
bands of zero transmission. In optical glasses these bands are located above 2200 nm in the
infrared (although A few extremely dense flints are transparent up to 4000 or 5000 nm) and
below 420 nm to 320 nm in the ultraviolet. "Red" and "blue" absorption causes window glass,
sometimes used for large corrector plates, to appear slightly greenish.
Because each wavelength of light is refracted at a slightly different angle by the same lens, the
index of refraction must be standardized on a specific wavelength of light, defined as one of the
emission lines of a chemical element. By convention this is either the d wavelength of helium
"orange yellow" [called yellow] light at 587 nm (nd) or the e wavelength of mercury "yellow
green" [called green] light at 546 nm (ne). Then the Abbe numbers Vd and Ve are calculated as:
Vd = [nd1]/[CF]
Ve = [ne1]/[C'F']
where the comparison wavelengths used (image, left) are the emission lines of hydrogen at C =
656 nm (H) and F = 486 nm (H) for d yellow, and the emission lines of cadmium at C' = 644
nm and F' = 480 nm for e green. Note that V numbers increase as the difference in refraction
between the red and blue wavelengths gets smaller.

The diagram (above) illustrates the relative effect of the refractive index and V number in two
equilateral prisms; the actual dispersion has been exaggerated in the diagram for clarity. At the
highest dispersion (lowest Abbe number), the indices of refraction for "red" and "blue"
wavelengths differ by less than 0.08. When the V number is large, the spread between the
"orange red" and "blue" indices of refraction is small, compared to the amount of mid spectrum
refraction. Note however that dispersion and refraction are roughly proportional: as the power of
refraction increases, so also does the amount of dispersion produced.

When glasses are charted using those


two metrics, they populate an area
between diffraction indices of about 1.4
to 2.0, and V numbers between about
90 (low dispersion) to 20 (high
dispersion). Although dispersion
increases as refraction increases, the
two attributes are not perfectly related;
so it is possible to choose glasses that
have different properties of refraction
and dispersion, which, in combination
with the shape and spacing of lenses,
can produce a wide variety of optical
systems.
Glasses on the left (dispersion of 55 or
more at refractive index below 1.60 and
50 or more at refractive index above
1.60) are called crowns, and glasses on
the right are flints. Curving across the
lower right of the diagram is the glass
line of flints produced by combining
crown glass with lead oxide. The
characteristics produced by different chemical additives are shown as different background
colors; these produce small differences in refractive or dispersive power denoted by terms such
as hard, soft, dense and light.
The dots indicate specific glasses available from Schott, one of many major glass suppliers. They
illustrate that the distribution of glasses is not even, and is strongly clustered around the
righthand diagonal of the diagram, indicating that refraction and dispersion are correlated. Most
of the variations occur within middle values of the indices of refraction and dispersion, straddling
the boundary between crowns and flints. This permits the manufacture of compound lenses
made of two or more glasses that produce the same refraction but different dispersions, or the
same dispersion but different refractions, which is necessary to eliminate chromatic aberration
in the focused image.
Traditionally, crowns are "hard" glasses with higher melting temperatures, made with small
quantities of potassium oxide, combined with oxides of other metals such as phosphorus, zinc,
lantahnum or barium. Crowns made with boric oxide (borosilicate glasses, including Pyrex
glass) or fused quartz have been especially important in the manufacture low expansion
telescope mirrors. Glasses marketed as ED (for extra low dispersion) are nominally crowns
(including fused fluorite or calcium fluoride, CaF2) with a V number above 80 and a partial
disperion far from the Abbe line.
Flints are "soft" glasses with lower melting temperatures, historically made with varying
quantities of lead oxide. One of the highest index flints (nd = 1.96) contains 18% silica and 82%
lead oxide by weight; some flint glasses used in military applications have even been made with
heavier metals such as thorium or uranium. Flints are especially susceptible to clouding under
prolonged contact with moisture, which leaches lead from the glass. For environmental reasons,
lead oxide is now frequently replaced with oxides of lanthanum, titanium or zirconium.
In the late 20th century, a few hard plastics (acrylics, polycarbonates, urethane monomers) have
been successfully used as optical materials in many applications, and lenses incorporating
diffractive (ribbed or grooved) surfaces have been used as well. Plastics can be injection molded
but are also relatively soft and easily scratched, and can deform under moderately high
temperatures, which makes optical coatings impractical. However certain plastics have replaced
Canadian balsam as cements commonly used to bond together the separate elements of a
compound lens.
Optical Coatings
As explained above, a small fraction of the light directed through an optical system is not
refracted by an optical surface but is reflected from it, and additional light is absorbed by the
material and scattered by internal reflections at the second (exit) surface.

In general, the transmission of a light ray through both surfaces of a single uncoated optical
element of refractive index ni and surrounded by air is:
T = 2ni / (ni2 + 1)
For an angle of incidence of between 0 to ~30, the two surfaces of a single optical glass
element in air transmit about 92% (ni = 1.5) to 81% (ni = 1.95) of unpolarized light, assuming
there is zero light absorption by the glass itself. This reflection obviously reduces image
illuminance and also often causes scatter, glare or ghosts within the image itself. In addition, the
transmission of a multielement optical system is no greater than the product of the separate
transmission values of all elements: two lenses of 92% and 81% transmittances combined as an
eyepiece yields a total transmission no greater than 75% (and possibly less). Obviously, methods
that can reduce surface reflections are highly desirable.
Optical coatings are very thin layers of material vacuum deposited on an optical element to
minimize or control the reflection or transmission properties of its surfaces. A variety of materials
used for this purpose include fluorides and oxides of various metals, but the most commonly
used material is magnesium fluoride (MgF2), which provides both a protective coating to a glass
surface and a refractive index that is optimal for controlling reflections.
Coatings manage unwanted reflections by virtue of their optical thickness, defined as the physical
thickness times the index of refraction. The material is deposited as a single layer approximately
1/4 or 1/2 wavelength thick, using vacuum processes monitored photoelectrically with
monochromatic light.
Light reflected by the internal surface of a 1/4 wavelength layer will be exactly 1/2 wavelength
out of phase with light reflected by the external layer, resulting in destructive (wave canceling)
interference of all reflected light. This effect is maximized when the refractive index of the layer
(in air) is equal to the square root of the refractive index of the glass. Magnesium fluoride (n =
1.38, n2 = 1.90) is optimal for very high index flint glasses; it is preferred for crowns as well due
to the durable and protective coating it provides to the optical surface.
Due to its fixed thickness and refractive index, the wave canceling effect of a single layer optical
coating becomes less effective at wavelengths longer or shorter than the design wavelength
(usually around 550nm to 560nm). This permits some reflected light at "red" and "blue"
wavelengths, which combine to produce the characteristic purple tint of a single layer coating.
In multicoated optics, three or more layers of different thicknesses can be designed to eliminate
reflected light within a bandpass matching the visible spectrum (~400nm to ~750nm). These
lenses generally have a much darker greenish surface tint or lack surface reflections entirely. The
"super multicoating" (SMC) used on Pentax optics consists of seven layers and reduces the
design wavelength reflection to 99.8%. Coatings of 50 layers or more have been designed to
produce a very narrow bandpass (for use as a filter); this bandpass can be shifted to longer or
shorter wavelengths simply by increasing or decreasing the thicknesses of all layers by a
constant factor.
The commercial designations coated or multicoated indicate that one or more (usually only the
external) air/glass surfaces are coated. Fully coated or fully multicoated indicates that all
air/glass surfaces (including those inside the instrument) have been coated.
Aluminum mirrors are also coated with very thin layers of magnesium fluoride or silicon
monoxide (SiO) to provide protection against oxidation and cleaning. Certain multilayer coatings
can also increase the reflectivity of an aluminum coating from 88% up to 99% at certain
wavelengths, but not more than ~95% as an average across the entire visible spectrum.
Further Reading
Astronomical Optics, Part 2: Telescope & Eyepiece Combined - the optics of astronomical telescopes and eyepieces, separately and
combined as a system.
Astronomical Optics, Part 3: Optical Aberrations - an in depth review of optical aberrations in astronomical optics.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of historically important eyepiece designs.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces, and results from my collection.
Modern Optical Engineering by Warren Smith - a classic, authoritative and clearly written survey of optical principles and applications.
Basic Optics and Optical Instruments by Fred A. Carson - a simplified but useful exposition of geometrical analysis.
Ray Tracing of Thin Lenses by Darryl Meister - Clear and well organized explanation of thin lens ray tracing.
Nature and Properties of Light by Linda Vandergriff - overview of modern optical techniques for the detection and manipulation of light.
Basic Geometrical Optics by Leno Pedrotti - the basics of light reflection and refraction and the use of simple optical elements, such as
mirrors, prisms, lenses, and fibers.
Basic Physical Optics by Leno Pedrotti - the phenomena of light wave interference, diffraction, and polarization.

Astronomical Optics
Part 2: Telescope & Eyepiece Combined

Telescope & Eyepiece Combined


Three Basic Telescope Functions
Stops, Pupils, Windows & Baffles
Focal Length & Field of View
Telescope Designs
Refractors
Refractor Design Principles
Reflectors
Newtonian Design Principles
Cassegrain Design Principles
Refractor/Reflector Hybrids
Field of Full Illumination
Eyepiece Design
Telescope Optical Attributes
Collimation
Focus
Light Grasp

Relative Aperture (Focal Ratio)


Exit Pupil
Resolution
Magnification
Magnification Selection
Apparent Field of View
True Field of View
Effect of Central Obstruction
Optics of the Eye
Individual Differences
Illuminance & Luminance
Luminance Adaptation
Light Grasp
Resolution
Contrast Sensitivity
True Field of View
Focus Diopters

This page introduces the optical principles necessary to understand the design and performance
of astronomical telescope systems the telescope and eyepiece used as a visual instrument
with the eye included as a third component. It is one of series: the previous page explained
basic optics; subsequent pages discuss eyepiece optical aberrations, eyepiece designs and
evaluating eyepieces.
Included at the end of each page is a list of Further Reading that identifies the sources used
here and background information available online.
Telescope & Eyepiece Combined
In this section I describe the basic functions, parts and terminology that characterize the
Keplerian telescope an "inverting" objective combined with an eyepiece. Diagrams show the
telescope objective schematically as a single refracting lens, but it may be a dioptric system
(comprising one or more lenses), a catoptric system (one or more mirrors), or a catadioptric
system (mirror and lens combination). The objective is combined with an eyepiece, used as a
magnifier to inspect the detailed content of the objective image.
Three Basic Telescope Functions
An astronomical telescope has three basic functions:
(1) Light grasp. Our ability to see very faint (low luminance) objects is limited by the area of
the pupil opening of the eye, which admits only a small amount of light. The telescope admits a
column of light whose cross sectional area is many times larger than the pupil, increasing the
total illuminance of the image. Implicit is the complementary function of beam compression,
which means that the light gathered by the large aperture is concentrated into a much smaller
diameter beam that can completely pass through the relatively tiny pupil of the eye.
(2) Angular resolution. Our ability to see objects that are small or at great distances is limited
by the eye's nominal angular resolution which can be from about one to over six
arcminutes. A telescope transmits the image of a distant object through a very wide pupil,
producing diffraction limited resolution on the order of one arcsecond. Implicit is the
complementary function of magnification of the telescopic image, which means its arcsecond
details are expanded to match the arcminute visual resolution. This magnified image remains
sufficiently illuminated due to the light grasp of the telescope.

(3) Pointing. The eye as a mobile receptor can look in any direction and remain fixed on any
single object. The telescope must be able to match this capability so that its optical axis can be
aligned in any specific direction and the alignment sustained for extended observation or long
exposure photography. For telescopes on the surface of the rotating earth, this implies tracking
through a moveable mounting and sidereal motors.
Note that light grasp and beam concentration can be entirely analyzed in terms of geometric
light rays, but angular resolution and magnification must be analyzed in terms of physical light
wavefronts.
Resolution and light grasp are both exponential functions of
aperture - as reciprocal aperture (1/D) and aperture area (D2)
- but light grasp increases more than resolution in larger
apertures. This leads to the invariant optical relationship
between light grasp and resolution shown in the diagram
(right). For any arbitrary range of apertures up to a maximum
feasible aperture (defined as 1.0), then 1/2 the total gains in
resolution are realized at 1/2 the maximum aperture, which
yields only 1/4 of the total feasible gains in light grasp. These
proportions apply no matter what aperture we choose as the
feasible maximum. Thus, if we (arbitrarily) take the practicable
limit of amateur telescopes as a 700 mm (28") Dobsonian
reflector, then 1/2 the potential gains in resolution are realized
with an aperture of 350 mm (14"), which delivers only 1/4 the
potential gains in light grasp.
This cost/benefit calculation divides the range of available aperture into two regimes. The benefit
in telescopes below the half aperture point is primarily in their resolution, and any increase in
aperture within this range increases resolution proportionately more than light grasp. Above the
half aperture point, increases in aperture deliver proportionately more light grasp than
resolution. As a result:
Double star, lunar and planetary visual astronomers generally prefer high optical quality
telescopes in the "resolution" half of the regime, despite the comparative lack of light grasp
which favors refractors and reflectors with apertures below 300 cm that are less sensitive to
atmospheric turbulence and permit a longer focal length (which produces greater objective
magnification)
Deep sky astronomers generally prefer high transmission and small focal ratio instruments in
the "light grasp" half of the regime above 300 cm despite the resolution problems that arise
from mirror flexure, miscollimation and atmospheric turbulence, and the field curvature and
coma/astigmatism produced by a small focal ratio.
Astrophotographers prefer telescopes at or just below the inflection point in aperture and at
moderate focal ratios, maximizing light grasp and field of view without introducing excessive
optical aberrations or sensitivity to atmospheric turbulence.
For visual use, a key feature of astronomical telescopes is that the combination of telescope
objective and eyepiece creates an afocal optical system. The objective focuses a beam of
parallel light rays from an "infinite" (far distant) object or object space (such as the celestial
sphere) as a point on the image plane at the apex of a converging light cone. The eyepiece
receives the light rays diverging from the opposite side of the image plane and focuses them a
second time into a much smaller exiting beam of collimated light rays.
As a result, the telescope and eyepiece combination does not have a single focal point or focal
length, although the separate objective and eyepiece do. This allows variation in true field, the
basis of resolution, and field width, the basis of luminance.
The afocal output provides a comfortable visual inspection of the telescope image. The eye is
structured so that small muscles can cause the eye lens to bulge, reducing its focal length to
focus the diverging light rays from nearby objects. Infinitely distant objects enter the eye as
parallel light rays, and to focus these the focal eye muscles become completely relaxed. Because
a telescope produces an exit beam of parallel light rays from every object point, the eye can
view the telescope image in a completely relaxed state, allowing extended viewing without
strain. When the telescope is used with a camera or other recording instrument, the afocal

output is not required and the telescope is typically used without an eyepiece.
These generic telescope functions light grasp and beam compression, resolution and
magnification, pointing and tracking may be adapted in different ways for solar astronomy,
spectroscopy, wide field photography, transit measurements or field portability, but all are
important in telescopes used for visual astronomy.
Stops, Pupils, Windows & Baffles
The light grasp and beam compression functions of a telescope are defined by stops, which are
physical obstructions within the system. The optical lenses and mirrors of the system create
images of these obstructions, called pupils and windows, which can be viewed from either the
object or observer end of the telescope (diagram, below). Baffles are used outside the path of
the light cone in order to minimize internal reflections.

The optical axis is the axis of rotation for all refracting or reflecting surfaces; mirrors, lenses,
stops and baffles are centered on the axis and perpendicular to it, with the exception of tilted
optics such diagonal mirrors or prisms.
Axial light rays are emitted from a point on the optical axis; if the point is sufficiently far away
(at optical infinity), the light rays reach the telescope as a collimated beam of rays parallel to
each other and to the optical axis.
A stop is a physical obstruction that limits the amount of light entering or leaving the telescope
system.
The aperture stop is the obstruction that limits axial light rays entering the telescope system:
its interior diameter defines the aperture of the telescope. In astronomical telescopes this is
either the mounting of the refracting lens or corrector plate, or the circumference of a reflecting
mirror. The aperture stop defines the principal point of the objective; its position affects the
appearance of aberrations in the image created at the objective focal surface.
The field stop is the obstruction that limits the area of the image or the amount of light exiting
the telescope system. It is usually coincident with the objective focal plane, and is either a barrel
diaphragm or field lens lock nut in the eyepiece, or the edges of a camera CCD chip or film
holder. The field stop controls the apparent field of view (the angle of marginal rays passing
through the eyepiece or onto the CCD sensor), and it limits the appearance in the image of
aberrations that increase with field height.
The linear diameter of the aperture stop defines the clear aperture (Do), which determines the
angular resolution of the objective. For reflecting telescopes with a central obstruction
(secondary mirror support), the area of the secondary obstruction (ds) must be subtracted from
the area of the aperture stop to compute the effective aperture (Deff) that determines the light
grasp of the system.
A pupil is the image of the aperture stop, viewed from a point on the optical axis at either end of
the telescope.

The entrance pupil is the virtual image of the telescope aperture stop, illuminated from the
image space and viewed from the telescope object end.
The exit pupil is the virtual image of the telescope aperture stop, illuminated from the object
space and viewed from the observer end. (The images of the aperture stop with or without the
eyepiece in place are both exit pupils, but standard practice is refer to the exit pupil produced by
the objective and eyepiece combination.)
In the UK the exit pupil is called the Ramsden disc, the specific plane cross section of the
telescope's afocal output where the image of the exit pupil has the smallest diameter. When the
observer's eye pupil coincides with the Ramsden disc, all the light collected and focused by the
telescope will enter the eye, provided the eye pupil is large enough.
Pupils are images of the the same telescope obstruction, so the ratio of the linear diameters of
the entrance and exit pupils indicate the system magnification (Mt) produced by the objective
and eyepiece combination.
A window is the image of the field stop, viewed from a point on the optical axis at either end of
the telescope.
The entrance window is the virtual image of the telescope field stop, illuminated from the
image space and viewed from the object end of the telescope.
The exit window is the virtual image of the eyepiece or camera field stop, illuminated from the
object space and viewed from the observer end.
Baffles are interior obstructions, constructed as diaphragms or tubes, that block stray light and
internal reflections from entering the image area and introducing glare or reducing contrast in
the image. Baffles do not affect the aperture, image dimensions or appearance of aberrations.
Because they are not stops, baffle edges are always just outside the circumference of the light
cone concentrated by the objective.
Focal Length & Field of View
The angular resolution and magnification functions of the telescope system, including the true
and apparent fields of view, are defined by the diameters and optical powers of the telescope
objective and eyepiece (diagram, below). Note that all linear or area measurements are
conventionally (and for calculation purposes conveniently) expressed in millimeters, and angles
are expressed in either radians or degrees.

Axial rays originate from a point on the optical axis; for objects at infinity they arrive as a
collimated beam of rays parallel to the optical axis and perpendicular to the entrance pupil, or as
a flat wavefront perpendicular to the optical axis and parallel to the entrance pupil. Oblique rays
are any ray with a field angle greater than zero.
The aperture stop fixes the location of the objective entrance pupil. This is a plane

perpendicular to the optical axis of the objective. The objective back principal plane is the
second principal plane of the objective (mirror or lens) independent of the location of the
entrance pupil.
The intersection of the entrance pupil with the optical axis is the principal point of the
objective. A principal ray or chief ray is any geometric ray that passes through this principal
point.
The objective focal length o is distance along the optical axis from the back principal plane to
the objective focal point.
The objective focal plane is a plane (more often, a slightly concave surface curved toward the
objective) that intersects the optical axis at the objective (mirror or lens) focal point.
The location of points in the object space and of conjugate points on the image plane are
measured within the telescope optical beam by a system of angles and corresponding heights:
The field angle of an object point is the visual angle, measured at the objective principal
point, between the optical axis and the principal ray from any point in the visual object space.
Field angles are used in place of the field height h to locate points in an inaccessible or infinitely
far object space, and are expressed as either the radians or tangent of the angle when the angle
is not large (less than ~20), and can be expressed as the angle itself when the angle and the
tangent of the angle are effectively equal (at angles less than ~1).
The aperture height y is the point at which any incident ray, regardless of its field angle,
intercepts the entrance pupil. It is measured as the distance of the incident point from the
optical axis on the plane of the entrance pupil.
The object height h is the distance of an object point from the optical axis as measured in the
physical object space.
The corresponding image height h' is the distance from the optical axis measured on the
objective image plane. Its location on the image plane is defined by means of the field angle as:
h' = tan()o or (for larger angles) h' = o/57.3.
These heights and angles are used to identify the points where a light ray intersects the
entrance pupil or image plane.
The focal angle u' is the angle of refraction or reflection between an image ray and the optical
axis, measured in the plane that contains both the ray and its corresponding principal ray. The
focal angle is determined by aperture height y and field angle , and the power of the objective.
The image height h' of the field stop is the image radius, and twice this amount is the image
diameter. The field angle of the most oblique principal ray that can pass the field stop is the
field radius of the image, and twice this angle is the objecive true field of view (TFOV), the
maximum angular width of sky imaged by the objective.
The eyepiece is two or more refracting lenses mounted as a unit that can be conveniently
installed into or removed from a position immediately behind the objective focal plane. The
eyepiece functions as a magnifier that expands the objective image so that it can be seen,
enlarged and in focus, by the observer's relaxed eye.
The distance from the eyepiece field stop (focal plane) forward to the eyepiece front principal
plane is the eyepiece focal length (e). The location of the eyepiece front principal plane is
calculated from the separate focal lengths of the eyepiece components.
The exit pupil (Ramsden disc in England) is the the smallest diameter area that passes all the
light transmitted by the eyepiece. The exit pupil contains the smallest and brightest image
produced by the eyepiece, and is the only point at which the entire eyepiece image area is in
view; at a farther distance the exit pupil functions as a stop that only permits a keyhole view of
a small area of the image.
The distance from the exit pupil to the last air boundary of the eyepiece is termed the eye
relief or eye distance (ER). Astigmatic eyeglass wearers generally prefer eyepieces with a large
eye relief.

When attached to a telescope, the eyepiece does not project a conjugate image point. Light
from any point in the object space exits the eyepiece as a bundle (or pencil) of parallel rays
gathered across the entire aperture. Because the rays are parallel, they are afocal: they do not
form a unique focal surface. The image may be projected onto a card held at any distance from
the eyepiece, and this distance determines the image dimensions.
Because the exit rays from the eyepiece are parallel from any specific point in the image field,
they create the illusion that the eyepiece image arrives from infinitely far away. In that case,
when the eyepiece field stop is coincident with the objective focal surface, and the eye pupil is
coincident with the exit pupil, the image will appear in focus when the eye is fully relaxed, just
as it is when viewing a far landscape or the night sky.
Telescope Designs
This section illustrates the most common optical formats found in astronomical telescopes.
These systems may be dioptric (comprising one or several lenses), catoptric (one or more
mirrors), or catadioptric (a combination of mirrors and lenses).
Refractors (Dioptric)
The refractor is the oldest telescope format, dating from the "spyglass" designs fabricated
c.1610 by Galileo. Early refractors provided good light transmission but were hampered by the
manufacturing difficulties of pouring large, chemically homogenous, strain free and bubble free
glass blanks, grinding them to the required precision, figuring lenses to eliminate optical
aberrations (chromatic aberrations in particular), and mounting the instrument so that it could
be conveniently pointed and could track moving stars. These four difficulties were not overcome
until the 19th century (e.g., in the Dorpat refractor designed and fabricated by Joseph von
Fraunhofer), which as a result became the great age of refractors. Since that time refractors
have played a minor role in astronomical research although they are still the instrument of
choice for many amateur astronomers.
The fundamental constraint on refractors is dimensional. Without mirrors, the focal length must
be straight from objective to eyepiece. Refractors built in the 17th and 18th centuries minimized
chromatic aberration with focal lengths of 7 to 10 meters (at relative apertures of /100 or
more) that were extremely cumbersome to mount. The objective can only be supported around
its perimeter, but the maximum size of lenses that will not deform under their own weight is
around 1 meter. (The largest operational refractors ever made are the 91 cm Lick Telescope and
the 102 cm Yerkes Telescope, both manufactured by Alvan Clark & Sons.) Even at their
"modern" relative aperture of /19, these telescopes are extremely long.
The early advantages of refractors were in the relative simplicity of figuring their spherical
surfaces and the high transmission of glass lenses compared to speculum mirrors. Once
chromatic aberration, spherical aberration and coma were minimized through the use of
achromatic objectives and sophisticated optical fabrication (including hand correction of lens
surfaces), refractors were manufactured during the 19th century to very high optical standards,
and opened the door to fundamental research in lunar, solar, planetary and double star
astronomy, as well as the earliest applications of photography in astrometrics and spectroscopy.
Ray tracing a refractor is a straightforward application of the principles outlined in the previous
section. The entrance pupil defines the front principal plane and the principal point for all chief
rays. The back principal plane emits all rays at the refracted angle. The focal length is positive
and the image is negative (reduced, inverted and reverted).

Refractor Design Principles


Refractor objectives can be constructed from one to four separate lenses or elements, each
potentially having distinct surface curvatures and made of a type of glass with different
refractive and dispersive attributes. In combination, these are used to minimize optical
aberrations.

Type

Achromats
Fraunhofer
achromat

Inventor (date)

J. von
Fraunhofer (1824)

Typical
Relative Aperture

/6 to /20

The Fraunhofer achromat consists of a flint biconvex


external lens and a negative crown internal lens,
separated by a thin air space. It was first used in the
240 mm Dorpat refractor made for double star
astronomer F.W. Struve.

Steinheil achromat Steinheil (1840)

/6 to /20

The Cassegrain provides a long focal length within a


short optical tube assembly. First described in theory by
Mersenne, it was by Laurent Cassegrain in 1680.
Usable field is small, and off axis aberrations limiting.
Its high magnification makes it especially suitable for
lunar, planetary and double star work.

Apochromats
apochromat

Steinheil (1840)

/6 to /20

The Cassegrain provides a long focal length within a


short optical tube assembly. First described in theory by
Mersenne, it was by Laurent Cassegrain in 1680.
Usable field is small, and off axis aberrations limiting.
Its high magnification makes it especially suitable for
lunar, planetary and double star work.

The dominant design feature of a refractor is its focal ratio. A larger focal ratio (longer focal

length at a constant aperture) increases the system magnification, and because it requires a
lower power objective it minimizes optical aberrations, including chromatic aberrations.
Classical achromats in apertures up to 280 mm are available today in focal ratios up to /15,
and 19th century refractors were manufactured at focal ratios of about /19, where the
refractor's visual performance is superb in solar, planetary and double star astronomy.
In the interest of a wider field of view (for photographic uses) and portability (shorter tube
length), most refractors manufactured today have much smaller focal ratios, usually from /8
down to /4. To manage the intrusion of optical aberrations at the necessary higher powers, a
third lens is used in an apochromatic design. These are often marketed as APO or APO triplet
designs, for example by TEC, Astro-Physics, Takahashi and Sky-Watcher.
Certain types of optical glass, denoted ED or extra low dispersion glass, allow greater accuracy
in the design of refractive surfaces that reduce both coma and chromatic aberration. Achromats
made with an ED glass element are sometimes marketed as ED achromats or semiapochromats,
for example by Sky-Watcher
Reflectors (Catoptric)
The reflecting telescope is also a 17th century conception, originating in two basic designs the
small prototype reflector of Isaac Newton (c.1680) and the conceptual designs of Mersenne and
Cassegrain, which were not actually built in any numbers until the 20th century.
The majority of reflecting telescopes consist of one primary mirror to gather and focus light and
one secondary mirror to direct the converging light cone to an observing location that does not
obstruct the aperture.
The fundamental constraints on Newtonian reflectors were mass and reflectivity. Mirrors were
made of polished speculum, an alloy of copper and tin, which could only achieve a reflectivity of
~65% or less. The speculum mirror, mirror support and tube of the largest 19th century
reflector, the Rosse 72" telescope, weighed 12 tons; two mirrors were made, so that one could
be repolished while the second was used in the telescope. The first large telescopes using glass
mirrors coated with a durable silver film were the Mt. Wilson 60" and 100" reflectors; these
demonstrated the superiority of the reflector in large apertures. Today all large research
telescopes, and the majority of personal telescopes of all sizes, are reflectors.
In addition to issues of mass and reflectivity, Cassegrain designs were constrained by the
difficulty of figuring matched paraboloid, hyperboloid or ellipical surfaces, an optical challenge
that was not practicably overcome until the 20th century.
Tilted Component Telescopes (TCTs) originate in the off-axis viewing position used by William
Herschel with his 18.8 inch /12.8 reflector, intended to omit the light loss that would be caused
by a speculum secondary mirror. Later designs, culminating in the Schiefspiegler design by
Anton Kutter, were similarly designed to eliminate the secondary mirror obstruction. These
designs are rarely constructed and not considered here.
Newtonian Design Principles
The Newtonian reflector is the most reductive telescope design, in the sense that there is only
one surface of positive optical power. (The secondary mirror is flat and therefore in the ideal
case contributes nothing to the quality of the final image.) This greatly reduces the optical
issues. The first requirement is that the front edge of the telescope tube should be located at the
primary focal point. In other words, the distance of the secondary mirror from the front edge of
the telescope tube should be equal to the distance from the secondary mirror to the focal
surface imaged through the focuser. The major uncorrected aberrations in a Newtonian reflector
are coma and field curvature, which become visually negligible at focal ratios above /10 but
pronounced at focal ratios below /6.
Usually an added optical component is mounted in front of the eyepiece or camera to minimize
these aberrations at low focal ratios. A commercial coma corrector is usually a pair of compound
lenses (e.g., a negative doublet in front of a positive doublet) mounted so that they can be
inserted into a focuser drawtube with the objective focus located in the space between the two
groups; the corrector can accept an eyepiece or camera adapter at the output end. The front
negative element introduces negative coma (comatic tail pointing toward the optical axis) that
largely but not entirely compensates for the positive coma produced by the primary mirror.
These also slightly increase the eye relief, reduce field curvature and astigmatism, and enlarge
the image by about 1.1 times.

Type

Newtonian
Newtonian

Inventor (date)

Corrector

Secondary
Shape

Typical
Relative Aperture

Primary
Shape

Newton (1680)

plane

/4-/12

+paraboloid

+ellipsoid

/4 to /12

+ellipsoid

hyperboloid

/4 to /12

+paraboloid

hyperboloid

/8 to /10

+hyperboloid

Due to its ease of manufacture and light grasp,


the Newtonian has become the major amateur
reflector design. Because the light is not folded
backward it requires a long optical tube
assembly, so that it rarely exceeds /12; but very
short relative apertures are used (often with
coma correcting optics) to exploit its light grasp.
If the focal point is at the same distance from the
mirror as the entrance pupil, the system is free
of astigmatism. At short focal lengths coma can
become significant and is often corrected with
optics placed before the eyepiece.

Cassegrain
Gregorian

James Gregory (1663)

The Gregorian design was first proposed in


theory by Marin Mersenne in 1636, then by
Gregory, although actual instruments was first
produced by James Short in the mid 18th
century. The Gregorian provides an erect image
by passing light from the primary mirror
through its focal point, inverting the image,
where it is reflected by the secondary, inverting
the image a second time. This requires a
somewhat longer optical tube assembly than the
classic Cassegrain, generally viewed as a
drawback. The effective focal length is derived
as the primary mirror focal length 1 multiplied
by the secondary magnification, calculated as
BFL/f. Its high magnification makes it especially
suitable for lunar, planetary and double star
work. Usable field is small, and off axis
aberrations limiting.

Classical
Cassegrain

Laurent Cassegrain
(1672)

The Cassegrain was first proposed by Mersenne


in 1636, then by Laurent Cassegrain in 1680. It
provides a long focal length within a short
optical tube assembly. Usable field is small, and
off axis aberrations limiting. The hyperboloid
secondary provides greater diverging power and
therefore a shorter focal length. Its high
magnification makes it especially suitable for
lunar, planetary and double star work, and
allows use of lower power, longer eye relief
eyepieces. Its short tube is more securely
balanced on a mounting and less affected by
wind or bumping.

Ritchey Chrtien

H. Chrtien (1922)

This design was proposed by Henri Chretien,


following concepts laid out in 1905 by Karl
Schwarzschild that eliminated the spherical
aberration and coma of previous reflector
designs. Chrtien's aplanatic instruments using
hyperboloid surfaces are usually modified in
modern telescopes so that the secondary
magnification is greater (the secondary
obstruction is smaller) and surfaces are closer
to paraboloid. The RC provides perfectly round
star images considered optimal for photographic
work and is the dominant design in very large
aperture telescopes. It has considerable
astigmatism at field angles greater than about
0.7 and the strongest field curvature of any
design, which inflates the diameter of star
images near the edge of the image field.

Dall-Kirkham

H. Dall (1928)

spherical

/12 to /20

+prolate ellipsoid

The Dall Kirkham has the advantage that the


mirrors are relatively easy to fabricate and test,
reducing costs and improving optical quality;
the spherical secondary also significantly
reduces the strict collimation requirements of a
classic Cassegrain or RC. However, the Dall
Kirkham has coma that is about 2 to 6 times that
in a classic Cassegrain and a small field that is
unsuitable for wide field photographic work; the
image quality deteriorates significantly as
secondary magnification increases. (The
Pressman Camichel design reverses the mirror
figures, using a spherical primary and ellipsoid
secondary, but has coma about 4 to 12 times
larger than a classic Cassegrain.)

Collimation tolerances become much more stringent in Newtonian telescopes with focal ratios
below /6, and collimation of a Newtonian telescope requires multiple adjustments to both
primary and secondary mirrors. The task is made easier by the use of a Cheshire eyepiece.
The normal procedure is to first center the circular outline of the secondary mirror within the
circular aperture of the focuser drawtube; then to center the image of the primary mirror within
the outline of the secondary mirror, and finally to center the image of the secondary mirror
within the outline of the primary mirror.
In smaller focal ratios the secondary mirror must be offset slightly from a centered position on
the optical axis. The offset must be both by a small distance away from the focuser, and by an
equal distance toward the primary mirror.
Cassegrain Design Principles
The Cassegrain format now dominates in large observatory and satellite instruments and is
increasingly popular in commercial instruments for private use. These instruments have great
flexibility, and can be specialized for visual astronomy or astrophotography.
In the Cassegrain design, light is focused by a concave (positive) primary mirror, and
intercepted by a convex (negative) secondary mirror placed close to and in front of the primary
focal point (diagram below, top). The secondary magnifies the primary light cone to a longer
focus and narrower focal angle sufficient to pass through a hole in the primary mirror. The
system effective focal length is then equivalent to extending this narrower light cone back
from the system focal point until the diameter of the light cone equals the diameter of the
entrance pupil (diagram below, bottom).

Three parameters are fixed arbitrarily at the outset in a Cassegrain design:


(1) the clear aperture D1 of the positive primary mirror, which determines the overall
dimensions of the optical tube assembly and the telescope's light grasp and resolution;
(2) the system effective focal length , which determines the telescope magnification, relative
aperture and the diameter of the useable image area; and
(3) the amount of back clearance b necessary for the mounting and adjustment of an eyepiece
focuser or camera adapter at the visual back a distance that must include the thickness of the
mirror glass, mirror cell and back plate (proportionately smaller in larger aperture telescopes).
In most designs these parameters have no limiting effect on the others given a few obvious
constraints e.g., the back clearance should not require a visual back opening larger than the
secondary mirror diameter, the secondary magnification should not be so low that it produces a
large secondary obstruction, etc. Aperture, effective focal length and back clearance are
determined solely by the usage requirements and construction details of the system.
The effective focal length is usually chosen to produce a relative aperture no less than Nt =
~12 and sometimes as much as Nt = 20 to 30. (Below Nt = 12 the magnification and contrast
potential of the system is wasted; above Nt = 30 the image area becomes unacceptably small.)
(Note however that commercial Schmidt Cassegrains are commonly designed at around Nt = 10
and Ritchey Chrtien Cassegrains as low as Nt = 8.)
The principal design decision is how to divide the necessary beam compression between the
primary focal length and secondary magnification. To optimize optical quality while minimizing
manufacturing costs and producing a sufficiently short optical tube assembly, the primary
relative aperture is usually chosen between N1 = 2 to 4. Then the required secondary
magnification (M2) is:
Nt = /D1 and N1 = 1 / D1
M2 = /1 = Nt /N1
Nt = N1M2.
The same system relative aperture is produced by increasing the primary focal length while
reducing the secondary magnification, or vice versa. The preference is for greater secondary
magnification, as this reduces the secondary obstruction from roughly 43% (at M2=2) to roughly
18% (M2=6). (A smaller secondary obstruction significantly improves image contrast for visual
astronomy but increases the difficulty of baffling unwanted light from the focal surface; a larger
obstruction is acceptable in instruments designed for astrophotography.) The drawbacks are that
higher magnification reduces the useful field of view and increases image aberrations and field
curvature.
Once M2 is fixed, and given the required value of b (in optimal designs b = D1 approximately),
the remaining dimensions (as labeled in the diagram) are:
f = (1+b)/(M2+1)
d = 1f
BFL = d+b
At this point it is useful to confirm both the minimum diameter of the secondary mirror that will
reflect all axial light in the primary light cone:
D2 = (D1BFL)/ = (D1f)/1
and the minimum diameter of the secondary mirror that will reflect all light (axial and oblique) in
the primary light cone:
D2 = [D1BFL/]+Ad/
where A is the diameter of the opening in the visual back. The secondary diameter should be
kept to a minimum consistent with the desired magnification, size of fully illuminated field and
effective baffling, usually around 25% of the clear aperture.
The remaining focal lengths, intermirror spacing and radii of curvature of the Cassegrain system
are:

= BFL+(M2d)
1 = d+(BFL/M2)
2 = (d+b)/(M21)
r1 = 21
r2 = 22 .
The field curvature () is:
1/ = 2/r12/r2
greater curvature arising predominantly from a shorter primary focal length. Both curvature and
coma become inconsequential for visual use, due to the small marginal image radius, calculated
as:
y = arctan(A/o).
However for photographic work a field flattener is often employed. This is usually a doublet lens
consisting of a field positive semiconvex lens cemented to a negative semiconvex lens. This
creates a strongly negative field curvature which counteracts the positive field curvature
inherent in the Cassegrain design with a small focal ratio primary mirror. Since a field flattener
will also contribute some astigmatism or coma, the field flattener must usually be designed to
work with a specific Cassegrain design and the amount of coma or astigmatism the system
produces. The recently introduced EdgeHD instruments from Celestron have a field flattener
built into the system, as do the Dall Kirkham designs from Planewave and the corrected Mewlon
(Dall Kirkham) designs from Takahashi.
For a field stop A = 32 mm (1.25 inches) in diameter and a 254 mm (10 inch) /20 system, the
objective image diameter is about 22 arcminutes.
Refractor/Reflector Hybrids (Catadioptrics)
Catadioptric systems originate in instruments designed c.1930 for wide field (survey)
astrophotography. They consist of a Newtonian or Cassegrain format reflector with one or more
refracting elements added. The refracting component is either one or more corrector plates
located at the aperture stop, or a compound lens located just before the prime focus as a coma
reducer or field flattener.
The purpose of the correctors in all cases is to improve the off axis optical quality of the
telescope, particularly coma and curvature of field, which is commonly desired in order to
shorten the focal ratio and increase the width of field. Thanks to modern methods of robotic
optical manufacturing, catadioptric designs are today one of the most popular formats for
amateur telescopes.

Type

Inventor (date)

Maksutov

D. Maksutov (1944)

The Maksutov was conceived as a robust, fully


closed optical system for use in elementary schools.
Amateurs rapidly adopted the design, and it was
early manufactured in the USA by Questar. The
design was optimized for /23 by John Gregory in
1957. The original design produced the negative
(convex) secondary as an aluminized spot on the
interior surface of the corrector plate; this reduced
the degrees of freedom needed to optimize the
aberrations of the system. The design by Harrie
Rutten (called a "Rumak") mounts a separately
figured secondary on the corrector at a longer
relative aperture.

Corrector
spherical
meniscus

Secondary
Shape
spherical

Typical
Relative Aperture
/4 to /12

Primary
Shape
+spherical

Schmidt
Cassegrain

D. Schmidt (1944)

aspheric
meniscus

hyperboloid

/4 to /12

+spherical

The Schmidt camera was developed to produce very


wide field images with no coma. The design was
adapted as a commercial telescope attractive for its
compactness and optical quality. The low power
aspheric lens at the aperture stop reduces spherical
aberration and provides support for the secondary
mirror mounting.

Catadioptric systems have become a dominant design, with refractors and Dobsonians, in
amateur (commercial) telescopes. All are modifications of either the Maksutov or Schmidt
camera designs, with Maksutovs dominating in the smaller apertures, including spotting scopes.
Both have the advantages of a closed optical tube assembly that minimizes instrument
convection turbulence and provides very short OTA dimensions and dust protection for the
primary mirror. They are relatively light weight and easily portable for the aperture, and unlike
Newtonian designs they do not require ladders or platforms for viewing near the zenith with
large apertures.
The optical design details of commercial Schmidt Cassegrain (SCT) designs are proprietary, but
several sources (including Rutten & van Venrooij and Smith, Ceragioli & Berry) have published
analytic deconstructions.
Field of Full Illumination
Not all the light incident on the objective contributes to the image at the objective focal surface.
Physical telescope construction means that some part of the off axis collimated rays passing
through the entrance pupil will be blocked from reaching the focal plane of the objective,
reducing image illuminance as field angle increases.
The limitation of field illumination as a function of field angle or field height is vignetting, and it
divides the image area into two parts: the central field of full illumination in which an object is
imaged with all the light that passes through the entrance pupil, and the concentric vignetted
periphery out to the field stop where the image illuminance is reduced as field angle increases.
While a dimming of the image at the field circumference of 50% or more typically goes
completely unnoticed in visual astronomy, it becomes a hazard in variable star astronomy
(where comparison stars in the same field are used to judge the variable star's magnitude) and
a blemish in nearly all photographic applications.
The field of full illumination is always calculated as a function of the field angle of the off axis
rays and the radius of a limiting field angle or vignetting obstruction internal to the telescope.
This can be either the desired widest angular width or linear dimension of full illumination
produced at the field stop (which determines the radius dimensions necessary for all secondary
mirrors, stops and baffles), or the secondary mirror minor axis (which controls the effective field
of full illumination), or the internal diameter of a refractor or Cassegrain stop or baffle (in
refractors, usually the last stop before the focal surface).
The metric in vignetting calculations is the maximum field angle of a bundle of off axis rays at
which no vignetting occurs. The diagram (above) summarizes the dimensions that are necessary
for the calculation in the three basic telescope formats.
(1) The first limiting angle on the field of full illumination is the shadow cast on the objective
diameter by the edge of the dew shield or tube opening (in a refractor or catoptric reflector) or
the dew shield, corrector plate or lens mounting (in the case of a catadioptric Maksutov
Newtonian, Maksutov Cassegrain or Schmidt Cassegrain):
tan(1) = 0.5(CDo) / t.
Most refractors are designed with retractable dew shields that vignette at field angles of 3 or
more. For a typical (well designed) Newtonian telescope, where the tube diameter is ~50mm
larger than the mirror (to minimize the effect of tube currents) and t = o (to minimize off axis
aberrations), this angle is 25/o or 0.7 in an /8 254mm system. For an /8 254 mm

Maksutov Newtonian with a corrector aperture of 300 mm the angle is 0.65. For an /10
254 mm Schmidt Cassegrain with an /2 primary mirror that is ~6mm larger than the corrector
aperture stop, the angle is 3/(2Dos) or 0.45.
(2) The second limitation (in a reflector) is the minor axis (ds) of the diagonal secondary mirror
in relation to the diameter of the light cone converging from the primary mirror. This in turn is a
function of the distance of the secondary mirror surface in front of the primary mirror focal
point, which in a Newtonian reflector depends on the primary mirror relative aperture, the tube
diameter and the desired "working distance" (w) outside the tube, and in a Cassegrain depends
on the secondary magnification. The generic formula is:
tan(2) = 0.5[ds (s/No)] / (os).

For an /8 254 mm Newtonian reflector with a 20% secondary obstruction, 25 mm tube


clearance and a 120 mm working distance (s = 272 mm), this angle is 0.5[51(272/8)] / 1760
or 0.28. For a Schmidt Cassegrain with a 30% central obstruction, an /2 primary mirror and
s =130 mm, this angle is 0.5[76(130/2)] / 378 or 0.83 in a 254 mm system.
(3) The third limitation is the internal diameter of the last smallest obstruction before the focal
surface, which may be the last tube baffle (in a refractor), the cylindrical or conical primary
mirror baffle (in a Cassegrain), the opening in the visual back, or the internal diameter of the
focuser drawtube (symbolized in the diagram above as opening z in relation to distance w):
tan(3) = 0.5[z (w/No)] / (ow).
For an /8 254 mm refractor or Newtonian reflector with a 120 mm working distance (w =
120 mm) and a drawtube internal diameter of 51 mm, this angle is 0.5[51(120/8)] / 1912 or
0.54. For an /10 254 mm Schmidt Cassegrain with a 42 mm central baffle opening at w =
360 mm, this angle is 0.5[42(360/10)] / 2180 or 0.08.
At each obstruction, for any angle greater than , the amount of the light reduction increases in

approximate proportion to the ratio cos()/pi. Vignetting effects can be calculated separately
and are linearly subtractive. In all cases, the very small foreshortening of the entrance pupil or
objective optical surface that occurs in off axis rays is negligible in comparison to the effect of
internal obstructions.
Refractors are vignetted only by the drawtube or last tube baffle which often has a diameter
close to the widest eyepiece field stop to minimize off axis aberrations. Newtonian reflectors are
constrained by the size of the secondary obstruction, which can be made larger to provide a
wider field of full illumination at the cost of slightly reduced image contrast. All Cassegrain
designs are significantly limited by the front opening of the primary baffle and in some cases by
the opening in the visual back; at large effective focal ratios (typically > 18), depending on
the diameter and magnification of the secondary mirror, no baffling solution may be possible
that blocks all stray light without vignetting. In no design is the tube opening the most
significant vignetting obstruction.
The calculations given above, although generic, illustrate the relative constraints on field size in
the different telescope designs:
design

tube

secondary

baffle

drawtube

Refractor
Newtonian
reflector
Maksutov
Newtonian
Schmidt
Cassegrain

>3.00

(0.54

0.54)

0.70

0.28

0.54

0.65

0.28

0.54

0.45

0.83

0.08

Newtonian Reflector. Because the principal constraint on the field illumination and image
contrast in a Newtonian reflecting telescope is the secondary mirror minor semiaxis (ds), the size
and position of the secondary becomes a critical issue in the telescope performance.
The formula for the proportion of full illumination (I) at linear field height h' is:
Ih' = [A/90] - [h'No(2/s2/o)sin(A)/pi]
where o is the effective focal length of the objective, N is the relative aperture of the objective,
and S is the distance from the secondary mirror (diagonal) to the focal plane. The angle A is
defined as:
cos(A) = h'No(1/S-1/o).
Cassegrain. Vignetting and baffling in a Cassegrain design are complicated by the fact that off
axis illumination or "stray light" can enter the focal plane directly, producing contrast loss and
ghost reflections. Eliminating this requires adjustment of three factors: the size of the secondary
mirror, the desired angular diameter of the image area (physical diameter of the field stop), and
the secondary magnification (distance from secondary to focal plane). The diagram (below)
illustrates some of the complications.

The dotted lines show the path of a column of collimated rays, from entrance pupil to
convergence at the focal plane; the lines are continued through the secondary mirror to show
the convergence of rays at the primary mirror focal point (1).
The yellow rays outline a bundle of rays fully illuminating the primary mirror that originate in a
source located off axis by an angle = 1. These are rays not intercepted by the front edge of
the telescope tube. They are reflected from the primary mirror and produce an angular
displacement from the collimated focal point. This displacement can be enough to cause some
part of the light cone to strike the secondary baffle or to miss the secondary mirror entirely. In
addition, the converging light cone is reflected from the secondary mirror at an angle typically
greater than 2 how much greater depends on the magnification of the secondary and this
can deflect the light cone so that a significant part of it is obstructed by the the primary baffle or
the central opening in the primary mirror.
Working from the other direction, the orange rays outline a light cone emerging from a point
inside the field stop that just avoid the primary baffle and strike the outer edge of the secondary
mirror. It turns out that these rays must be reflected from points inside the full aperture of the
primary mirror (reducing the effective aperture), and in addition are limited to a field angle of
less than 0.4. Thus the diameters of the primary baffle and secondary mirror limit both the
usable angular field and the aperture of the primary mirror.
Finally, the green ray shows that these restrictions on collimated rays are still insufficient to
eliminate off axis illumination or "stray light", which is not blocked by either the secondary
mirror or the primary baffle and can reach the image area inside the field stop. Stray light
illuminates the image area, reducing image contrast and causing a variety of ghost reflections.
To eliminate stray light the secondary obstruction must be made a larger proprtion of the
primary aperture, the primary baffle must be extended further toward the secondary mirror, or
the usable field must be made smaller. The diagram shows that extending the baffle will block
the orange rays, further reducing the area of the primary mirror that can illuminate the off axis
areas of the image; enlarging the secondary obstruction reduces the contrast in the image.
The standard remedy is to increase the diameter of the secondary mirror and opening in the
visual back, which increases the angular diameter of the usable image area at the cost of a
slight reduction in image contrast. Rutten & van Venrooij also suggest a conical shape in the
primary and secondary baffles, so that the flared edges of the secondary baffle will block light
that might be admitted by the primary baffle. However the secondary mirror diameter can be
minimzed and create the possibility of stray light, provided that a small angular field is
acceptable, planets are viewed only on axis, and stray light from the Moon is accepted as
unavoidable.
Eyepiece Design
Modern eyepieces typically consist of the few components, shown in the cutaway diagram
(below) for the 28mm RKE eyepiece used as the example of eyepiece optical design. This
illustrates the conventional eyepiece terminology and basic features.

First, an eyepiece consists of a few simple mechanical components, shown above with
underlined labels.
The mount is a turned metal or delrin cylinder with a threaded and centered cylindrical interior
chamber, the lens cell. Lenses are placed inside the cell, separated from each other as necessary
by metal or plastic spacer rings and secured at the open end by a threaded lock ring that
requires a spanner wrench to tighten or remove.
The mount ends on the telescope side as a flat shoulder which fits snugly against the eyepiece
holder and aligns the eyepiece to the optical axis.
The mount exterior is typically embossed with the eyepiece type, focal length (in millimeters)
and manufacturer's brand, and may be fitted with distinctive but basically decorative grip rings
or plastic housings.
The barrel is a threaded metal tube usually made of nickel plated steel, brass or aluminum,
used to secure the eyepiece in the focuser or eyepiece holder. The barrels of all modern
astronomical eyepieces manufactured in the United States have either a 1.25 inch (31.5 mm) or
2 inch (50.5 mm) nominal exterior diameter; many older eyepieces some eyepieces
manufactured in Japan have a 0.965 inch (24.5 mm) nominal diameter.
The interior surfaces of the barrel should be blackened (or made of black delrin or anodized
metal) to minimize scattered light. High quality eyepieces are also manufactered with internal
threads at the interior field end so that threaded filters can be mounted in the eyepiece.
The field stop is a rigid diaphragm or beveled ring fixed on the interior of the eyepiece barrel.
In shorter focal length eyepieces it may be defined by the interior edge of the field lens lock
ring. It is located at the eyepiece focal plane and provides both a crisp boundary to the field of
view and a mask to eliminate optical aberrations or vignetted image areas at wide field angles.
It can also be used to occlude a bright star or planet so that a faint companion or satellite is
easier to see.
Positive eyepieces place the focal surface and field stop in front of the field lens, outside the lens
assembly. Negative eyepieces (such as the Huygenian) place the focal surface somewhere
between the field lens and eye lens.
The undercut is a continuous flat notch in the eyepiece barrel, located to fit over the
compression ring of the telescope eyepiece drawtube. It is designed to prevent the eyepiece
from falling out of the focuser when the telescope orientation places the eyepiece in a downward
direction. Undercuts are a necessity in modern wide angle or super wide angle eyepieces, which
in the longest focal lengths can weigh 2 or more pounds; but in the traditional standard field
eyepieces that weigh only a few ounces undercuts are an annoying superfluity.
The optical components of the eyepiece include the following:
A lens is a single piece of glass; the component lenses of an eyepiece are called elements.
(The eyepiece in the diagram is a three element design.) A lens group or compound lens is two
or more lenses cemented together with balsam or resin; the compound lens is then considered a
single element.
The lens or compound lens at the telescope end of the eyepiece is the field lens; the lens at
the observer's end is the eye lens. Lenses in the middle of the eyepiece have no designation.
The opening in the field stop that limits the image to be magnified by the eyepiece is the
entrance window. The smallest area that contains all the focused light from the eyepiece is the
exit pupil (de). The eye receives the widest, brightest and least distorted view of the telescope
image when the eye pupil and exit pupil are exactly coincident and concentric. The exit pupil can
be viewed and measured as a small circle of light that appears just above the eye lens when the

telescope is turned toward the daytime sky.


The focal length (EFL) of the eyepiece is measured from the eyepiece focal plane to the
eyepiece principal plane, and is typically stated in millimeters.
The eyepiece field radius () is the angle between the optical axis and the image of the field
stop, as measured from the eye point (the intersection of the exit pupil and optical axis). Twice
the field angle (2) is the apparent field of view
(AFOV) produced by the eyepiece.
The eye relief is the available distance between
the outer surface of the eye lens and the pupil of
the observer's eye. This distance must comfortably
accommodate the space requirements of the
cornea of the eye, the observer's eyelashes,
contact lenses or eyeglasses (necessary to correct
astigmatism in the eye). The actual eye relief can
also be reduced by a bulky or protruding eyepiece
mount.
Some observers will fit eyepieces (especially those
with large eye relief) with a latex or plastic eye
cup or eye guard that provides support for the
observer's eye (or eyeglasses, if those are worn while observing), steadies the alignment and
distance of the eye to the eyepiece, and eliminates stray light (image, right). Many high quality
eyepieces come equipped with adjustable eye guards.
Lenses are usually modified with optical coatings to minimize reflections from the lens
surfaces. Eyepieces are multicoated if two or more coating layers have been applied to the lens
surfaces; they are fully coated if all lens surfaces, including cemented surfaces, have at least
one coating.
Telescope Optical Attributes
Collimation
Collimation is the condition in which the three optical axes of the objective, eyepiece and
observing eye are exactly aligned and coincident. It is the first of two critical adjustments
necessary to obtain the peak optical performance from a telescope. Miscollimation arises when
any of the three optical axes is out of alignment with the other two (diagram, below).

Observing through a spherical symmetrical lens from an off axis position has the same optical
effect as observing on axis through a distorted or asymmetrical lens. Miscollimation therefore
produces aberrations in "on axis" or centered star images where optical images are usually
flawless. These aberrations most often resemble coma or astigmatism. They are easiest to
produce by moving your head from side to side while viewing a star centered in the field of a
wide angle eyepiece or in a small focal ratio objective.
Objective Collimation. Collimation usually defined as the procedure of aligning the optical axes of

the objective lenses or mirrors with the center of the eyepiece focuser. This is the most common
source of mechanical miscollimation and also the type that is easiest to correct.
Adjustment knobs or screws are usually provided for this purpose in the mirror cells of the
primary and secondary mirrors of a reflecting telescope. It is not normally possible to collimate
the objective lenses of a refractor, or the corrector plate of a catadioptric (Schmidt Cassegrain,
Maksutov Cassegrain) telescope.
The specific steps necessary to collimate a telescope vary with the telescope design and
manufacturer. Nearly all commercially available telescopes are provided with illustrated
collimation instructions, and instructions are available online.
A number of collimation aids are commercially available to assist the collimation procedure.
The oldest and simplest is the Cheshire eyepiece (image, right) which is an open metal tube
supporting centered crosshairs and capped at the observer end with a pinhole. The crosshairs
are illuminated by a side opening, which is turned toward the sky or an observatory light. The
"eyepiece" is placed in the focuser, and the crosshairs viewed through the pinhole; their
intersection is intended to show the direction of the optical axis created by the focuser
drawtube. The secondary mirror is aligned with the focuser when it is centered on the
crosshairs; then the primary mirror is adjusted until the secondary mirror appears centered
within it.
Laser collimation tools are available to assist collimation, and these are especially popular
among the owners of large aperture, small focal ratio Dobsonian telescopes, where exact
collimation is required and tolerances can be quite small. Current vendors include (among
many) Baader Planetarium, Farpoint, HoTech, Howie Glatter and Catseye Collimation.
These tools are generally effective only with Newtonian format telescopes (including Maksutov
Newtonians), not Cassegrain format telescopes. HoTech offers a system that is specifically
designed for Cassegrain telescopes, but it is expensive ($400 to $450 at this writing).
However, visual collimation is highly effective and entirely adequate in moderate to large focal
ratio (/7 or more) telescopes, and the principles are the same in all cases where a star image
is used as the collimation target. The procedure is as follows:
(1) First a "rough"
collimation is performed
according to the
manufacturer's directions
so that the secondary
mirror appears centered
in the focuser opening
and the primary mirror
appears centered in the
secondary mirror.
(2) A moderately bright
star (magnitude 2 to 4)
at a medium zenith
angle (20 to 40) is
chosen as a target. In this orientation, the adjustment screws in either the primary or secondary
mirror cells must be easily accessible. The star is centered in the field of a low power eyepiece
and brought into defocus, with the eyepiece focal plane beyond (outside) the objective focal
plane, in the extrafocal position. (SCT users without an eyepiece focuser should note the SCT
focusing procedure, described in the next section.)
(3) The adjustment screws are used to center the secondary shadow within the circular star
image. The choice of adjustments depends on whether the primary or secondary mirror is being
moved and whether the star image is in extrafocal or intrafocal defocus. In all cases, the
defocused star image will appear as a thick annulus or nested rings of light around the
secondary shadow or a defocused star image. The rings will not be concentric, but crowded
together on one side of the star image (diagram right, top #1), and the goal is to manipulate
the star image until the rings are perfectly concentric (#2).

The trick is to think of the rings as forming a cone, whose axis of rotation is the optical axis
(diagram right, bottom). The cone is tilted to one side, which brings the rings on the opposite
side closer together. The adjustment knobs are used to shift the star image in the direction
away from this "crowded" or brighter side of the star image. This causes the star image to move
out of the center of the eyepiece field, and it must be recentered to evaluate the direction and
amount of the next adjustment.

(4) When the low power collimation appears reasonably accurate, a medium power eyepiece is
used to examine the defoused image and further adjustments are made, this time with a smaller
(less defocused) star image.
(5) The medium power eyepiece is replaced by a high powered eyepiece, and collimation is
adjusted further. At this point the out of focus star image will be so small that the secondary
shadow is no longer visible. Instead the Poisson spot that appears at the center of the
extrafocal star image is centered inside the star image circumference (diagram, left). This is
often disrupted by atmospheric turbulence, but the concentric diffraction rings around the
Poisson spot can be used to judge the collimation and determine the necessary adjustments.
(6) Finally, the focused image of the brightest available star is examined, with attention paid to
the contraction of the image in the last millimeter of focusing, and to the tiny threadlike
granules that appear in the flare or glare around the star image. These granules should appear
to radiate away from the star image equally on all sides. Note that they will be quite sensitive to
the location of the observer's eye pupil, which must be centered on the exit pupil. (This occurs
when the eyepiece field edge is clearly visible on all sides.) If the eye pupil is centered, then the
shimmering filaments can be used to judge the last refinements in the collimation.
(6) Most primary mirror cells are equipped with lock screws, which should be gently and
gradually tightened in a repeated circular sequence until all lock screws are fixed but not firm
(only contact and not pressure between lock screw and mirror cell is necessary). Tightening one
screw all the way before moving to the next can disrupt the collimation.
Eyepiece Miscollimation. Miscollimation in the eyepiece is usually a more difficult problem to
resolve. It can arise because the eyepiece is poorly manufactured (a very rare problem), the
focuser is not centered on or perpendicular to the optical axis, or the clamping mechanism in the
focuser drawtube is not symmetrical.
The alignment of the focuser can be checked with a Cheshire eyepiece, and most high end
focusers have adjustment screws as part of the tube mounting plate. In cases of extreme
misalignment, the mounting screw holes may need to be enlarged or redrilled and the focuser
remounted on the telescope tube.
The clamping mechanism should always be used with eyepieces, despite the possible annoyance
of the undercut, to maximize alignment with the drawtube and minimize variation in the
position. The eyepiece should first be seated so that the shoulder is in contact with the rim of
the drawtube. The clamp should consist of a thumbscrew that exerts pressure on a flexible metal
band, seated inside an interior groove. To ensure even distribution of pressure, the long end of
an allen wrench can be used to rotate the flexible metal band in its groove so that the gap in the
band is located directly opposite the thumbscrew.

Alignment issues are exacerbated when a mirror diagonal is used. In a test of six mirror
diagonals available to me I found errors of up to 30 arcminutes in the alignment of the mirror to
the optical axis, and to my annoyance I discovered that the most expensive mirror diagonals
were not the most accurate. (One was shimmed with a putty adhesive, the other with black
tape.)
Eye Misalignment. Correct positioning of the eye pupil around the exit pupil is a third source of
collimation error. It is very easy to produce aberrations in the appearance of a perfectly focused
star simply by moving the eye away from its optimal position.
With this simple maneuver, the observer can identify the optimal location of the eye. It is a
matter of physical skill to find and hold this position while observing, but various supports
(chairs or ladders) can make this task enormously easier. Positions that require bending or
leaning, or strain in the neck, back or legs should be avoided completely.

The location of the eye pupil and orientation of the eye optical axis are also important (diagram,
above). The eye position is usually correct if the entire circumference of the field stop is clearly
visible, and if any aberrations are apparent only near the edge of the eyepiece field of view and
are also not more pronounced on one side of the field than the other. A longitudinal
misalignment of the eye, with the pupil in front of or behing the exit pupil, vignettes the field
with the eye lens mounting, which will appear out of focus, and also creates a "keyhole"
narrowing of the field. A lateral misalignment causes a curved edge "shutter" or obscuration to
intrude into the field, and a misalignment of the optical axis of the eye and telescope system
often causes the field stop opposite the direction of view to disappear, or creates a floating
lenticular or "kidney bean" shadow in the opposite side of the field.
Focus
Focus is the second of two critical adjustments necessary to obtain the peak optical performance
from a telescope. By design both the objective and eyepiece have a specific focal surface; focus
is the distance adjustment along the optical axis that brings the two into coincident position.
The objective focal surface is the point of reference:
The area on the objective side of the objective focal plane is called intrafocal or "inside the
focus". Moving the eyepiece out of focus in this direction causes the output rays to converge,
corresponding to an increasing positive diopter setting that is appropriate for eyes that
undercorrect (are farsighted).
The area on the eyepiece side of the objective focal plane is called extrafocal or "outside the
focus". Moving the eyepiece out of focus in this direction causes the output rays to diverge,
corresponding to an increasing negative diopter setting that is appropriate for eyes that
overcorrect (are nearsighted).

Focusing Mechanisms. In a typical rack and pinion (gear and linear toothed track) focuser or
Crayford (roller and cylinder) focuser, the eyepiece is adjusted by a pair of knobs on a
transverse shaft below the optical axis. The rotation of the knobs necessary to move the
eyepiece in or out is immediately obvious from the visible movement of the eyepiece. Rotating
the right hand knob clockwise moves the eyepiece toward the objective on the optical axis, in
the intrafocal direction; counterclockwise rotation moves the eyepiece away from the objective.

In a commercial Schmidt Cassegrain system, in which the focus is adjusted by moving the
hidden primary mirror using a mirror cell focusing knob protruding from the back of the
telescope, the necessary adjustment is less obvious (diagram, right):
Turning the mirror cell focusing knob counterclockwise moves the telescope focal surface
toward the observer, so that an eyepiece is moved out of focus in the intrafocal direction, and
an eyepiece out of focus on the extrafocal side is brought into focus.
Turning the mirror cell focusing knob clockwise moves the telescope focal surface away from
the observer, so that an eyepiece is moved out of focus in the extrafocal direction, and an
eyepiece out of focus on the intrafocal side is brought into focus.
Focusing Procedure. The optimal focusing procedure is always to start focus from an
extrafocal position the eyepiece focal plane must be in front of the objective focal plane
(diagram right, bottom). The reason has to do with how the eye focuses the diverging light
rays from nearby objects. In the relaxed used to view distant objects the eye lens is actually
stretched into a flattened figure by zonule fibers or ligaments anchored in the iris muscle. To
focus on nearby objects, the iris muscle contracts in a way that relaxes the tension on these
ligaments and allows the eye lens to bulge into a shorter focal ratio. Age hardens the lens,
reducing and (around age 60) eliminating the elastic ability of the lens to respond to these
changes in tension.
When an image is brought into focus from an intrafocal position, it approaches focus with a
"nearsighted" convergence of rays. The eye handles these as it would an object near the eyes,
and tightens the iris muscles in an attempt to focus; some of this tension remains even after
the image is brought into focus. In contrast, an image brought into focus from the extrafocal
direction approaches focus with a "farsighted" divergence of rays. The eye cannot
accommodate toward the image any farther than a completely relaxed state, and this remains
once the image is brought into focus. If you overshoot the correct focus and err onto the
intrafocal side, return to the extrafocal position and start again.

In commercial Schmidt Cassegrain telescopes, changes in the primary mirror alignment ("mirror
shift" or "mirror flop") occur when focusing with the mirror adjustment knob. The recommended
focusing procedure is first to turn the focusing knob clockwise to bring the eyepiece into an
extrafocal defocus, then to bring the eyepiece back into focus by a counterclockwise adjustment
in the intrafocal direction. This should be the same focusing procedure used during collimation
(except that the star is left somewhat out of focus on the extrafocal side), so that the mirror tilt
during collimation and in routine observation is the same.
The ability to produce a sharp focus is partly dependent on the optical quality and collimation of
the instrument and partly dependent on the amount of atmospheric turbulence. Nevertheless,
the "best" focus will always produce the visually smallest or most compact image of a star, and
the sharpest edge in a planetary disk or lunar terminator, when the target is placed at the
center of the eyepiece field of view, and will make visible the greatest number of faint stars in
the field. Under optimal adjustment and viewing conditions the image will appear to "snap"
unambiguously into the best focus although the "snap" is most vivid in small focal ratio
telescopes, and atmospheric turbulence can make this test unreliable.
When poor seeing makes focusing difficult, best focus is found by bracketing the focus. First
bring the star into intrafocal defocus to a recognizable disk visual diameter, then take the star in
the opposite direction, through the focal position into extrafocal defocus, until the disk reaches
the same visual diameter. To finish, rotate the focusing knob back by half the travel between the
two defocus positions.
Poor seeing can cause large variations in focus. Resist the temptation to continually adjust focus
in order to "chase the seeing". The best procedure is not to make continual adjustments, but to
find the focus that produces the most frequent intervals of focus. Patience, rather than
manipulation, is the bast focusing procedure: focus should be changed by very small increments
so that the quality of the new focus can be observed for several seconds.
High power focusing is assisted by the use of a moderately bright star, as the diffraction pattern
(or turbulence speckles, under poor seeing) will appear compact and clear at best focus.
Depth of focus is the distance in front of and behind the optimal focal point within which image
quality is still acceptably focused. If the standard of image quality is that it is not degraded by
more than 1/4 wavelength of light (the Rayleigh criterion), then:
DF = 2No2
Depth of focus is proportional to the square of the relative aperture. Thus an /20 Cassegrain
telescope has a depth of focus of 0.44 mm at = 550 nm, but an /4 Newtonian has only
0.02 mm. These tolerances are especially critical for astrophotographers, who typically use
automated (software driven) focusers to find the optimal focal point at the camera image plane.
Light Grasp
The light gathering power of the telescope is a function of its aperture area (Ao): as the aperture
increases, more light enters the system and is concentrated into the usable image. For a
refracting telescope this area is usually calculated as:
Ao = Do2
Because the secondary mirror in a reflecting telescope blocks a certain amount of light from
reaching the primary mirror, its area (ds2) must be subtracted from the area of the objective in
order to get the effective aperture (Deff):
Deff = sqrt(Do2 ds2)
Ao = Deff2
Thus, a 250 mm (10") telescope with a 50 mm (2") diagonal diameter has an effective aperture
of 9.6" or 245mm.
The relative light grasp of any two apertures is proportional to the square of their areas, and the
increase in light grasp with aperture scales as the square of the aperture ratio:
Ao = (D1/D2)2
Thus, compared to a 250 mm (10") objective, a 400 mm (16") objective gathers (400/250)2 =
1.62 = 2.56 times more light.

To anchor these relationships as measures of illuminance or brightness, the light gathering


power of the telescope can be expressed in three ways: (1) as the increase in light gathered
over the light that enters an unaided, dark adapted eye; (2) as the limiting magnitude of stars
visible with averted vision, both visual parameters; and (3) as the illuminance of the telescopic
image on a film or CCD sensor, a photographic parameter.
(1) Telescopic Gain Over Naked Eye Brightness.
The first criterion requires an estimate of the dark adapted pupil aperture. To measure: obtain a
complete Allen wrench set. At night, once your eyes are fully dark adapted (roughly 1/2 hour
after the last bright light exposure), close one eye and pass an Allen wrench back and forth in
front of the other eye. Hold the long shaft of the wrench vertically with the short end pointed in
the direction you are looking. If the wrench fully occludes a bright star, causing it to "blink",
then the wrench is larger than your pupil opening. Try different wrenches until you discover the
largest wrench that dims but does not completely block the star image. Then the width of that
wrench is the diameter of your dark adapted pupil ().
Then the measure of a telescope's light grasp in comparison to your naked eye is the ratio of the
two areas:
Lo = (Deff/)2t
where t is an estimate of the system's light transmission, which is 0.98 for each surface of
aluminum high transmission catoptrics, 0.90 for aluminum standard catoptrics, and
approximately 2% per centimeter of fully multicoated refractor or eyepiece glass. My dark
adapted pupil is 5.9 mm wide, so a 10" reflecting telescope with a 2" diagonal will gather at
most (245/5.9)2 x 0.98 x 0.98 = 1656 times more light than my dark adapted eye, not
accounting for light absorption by the eyepiece.
(2) Telescope Limiting Magnitude.
Because visual magnitude differences are a direct measure of luminance ratios (expressed in
logarithms), the telescope/eye ratio implies a 10" telescope limiting magnitude that is 1656
times fainter than the visual limiting magnitude.
Using the standard magnitude ratio formula (where b' and m' refer to the fainter luminance and
magnitude, respectively), and assuming a visual limiting magnitude of 6.5 yields:
log(b/b') = 0.4(m' m)
log(1656)/0.4 + 6.5 = m'
or a 10" telescope limiting magnitude of 14.55.
Magnification is a critical qualification in estimates of the visual brightness of a telescopic image:
the objective gathers light, but linear magnification spreads a fixed density of light over a square
visual area. Thus the ratio of telescopic to visual brightness becomes:
Lo = (Deff2/Mt22)t
This roundabout calculation has been made more convenient, and aligned with various observer
tests, as the following standard formula for a telescope's limiting magnitude:
m'o = 2.7 + 5log(Deff)
where Deff is in millimeters. This formula assumes a pupil aperture of 7 mm and a naked eye
limiting magnitude of 6.5. It predicts a limiting visual magnitude of m'o = 14.65 in the 9.6"
(245 mm) telescope effective aperture.
Attempts have been made to refine this straightforward calculation of limiting telescopic
magnitude by including the effects of more variables magnification, sky brightness, altitude,
light absorption, optical quirks of the eye always with the statistically dubious assumption
that a greater number of inputs will increase accuracy.
The graph shows estimated light grasp for stars when viewed with averted vision, calculated by
Bradley Schaefer. This model includes the effects of observer age, sky brightness and pupil
size, among other factors, and shows the impact of different magnifications. These predicted
values correspond reasonably well to the actual limiting magnitudes reported by over 300
observers. It also predicts m'o = 14.65 for a 10" aperture at 100x magnification.

In the graph, the blue curves for different magnifications assume a sky brightness that permits a visual
limiting magnitude of 6.0; the single green curve shows the effect of reducing the visual limit to 4.0. The
size of the dark sky increase in a 10" objective at 100x (m = 1.25) is roughly twice as large as light
grasp increase between a 10" and 16" objective! This underscores the potential benefit of a medium
aperture but compact and portable field telescope for observers with access to dark sky sites.
Subsequent and more intricate proposals by Nils Olof Carlin and Chris Lord do not improve the validity of the
methods by adding more variables, because the error in the actual observed magnitude across
observers, due to factors still omitted or impractical to measure, remains on the order of 0.5 magnitude
or more the difference between a 6" and 8" telescope.
To understand why predicting limiting magnitude is so uncertain, consider the variables that can affect it:
the observer's age, optical quality of the eye, eye transmission (including corneal and lens yellowing, and
macular pigmentation), optical quality of eyeglasses or contact lenses (if worn), pupil diameter, blood
oxygen level, cold or wind stress, eye fatigue, physical fatigue, posture or seating comfort, dark
adaptation, focused attention and observing experience; the telescope's light transmission (including
both objective and eyepiece, including dirt and dew), collimation, magnification, quality of baffling or
internal flocking, focus and Strehl ratio; the altitude and geographic location (especially latitude) of the
observing site; the altitude, azimuth and spectral class (color index), and the field height within the
image, of the stars being observed; the sky transparency (humidity, aerosols and particulates in the air)
and sky brightness (light pollution, including presence of the moon and planets); the atmospheric
turbulence or seeing; and thermal currents in the telescope itself.
Schaefer's final comment "the cause of the 0.5 magnitude uncertainty in the model ... remains
unknown" is hardly sensible. Even setting aside the many instrument and environmental variables,
estimates of visual limiting magnitude are an exercise in predictive psychophysics, which are inherently
imprecise and unreliable due to the very large differences in visual capabilities across individual
observers.
(3) Photographic illuminance.
The photographic illuminance of the telescopic image increases with aperture and decreases with relative
aperture, making it proportional to:
Io = (Do/No)2
Magnification is not a factor, because the sensor is utilizing the prime focus of the telescope.
Relative Aperture (Focal Ratio)
The relative aperture (No) of a telescope is its focal length divided by the diameter of its entrance pupil
or clear aperture (Do):
No = o/Do
It is also known as the focal ratio or ratio. The name and notation invites erroneous interpretation as a
fraction (/2 is "larger" than /6), although relative aperture is actually the quotient of a ratio. Thus a

relative aperture of /2 is comparatively smaller or photographically faster than /6, which is larger or
slower.
Relative aperture is not a consistent measure of any physical quantity other than the maximum field
radius produced by a telescope objective:
= arctan(1/No).

Relative aperture is sometimes referenced in the context of optical aberrations. Thus, coma is considered
to be a significant aberration in Newtonian reflectors shorter than /5, and the spherical aberration in a
spherical mirror or lens becomes negligible (by the Rayleigh criterion) when
N 3.4(Dcm)1/3 or 4.6(Din)1/3
Both depth of focus (defocus tolerance) and depth of field increase with relative aperture.
In photometric contexts, relative aperture expresses the proportion of unit illumination (aperture area)
per unit magnification (focal length) produced by the telescope objective. However, these units are
dimensionless, and comparatively useful only by holding one element constant. For example, as the
relative aperture gets larger for a constant aperture, the quantity of light projected onto a unit area of
the image plane decreases because the image is magnified. Equivalently, as the relative aperture gets
larger for a constant focal length, the quantity of light projected into the unit image area decreases
because a smaller aperture admits less light.
Relative aperture determines the ratio between eyepiece focal length and exit pupil diameter, independent
of aperture or focal length. Thus, an /10 telescope will produce an exit pupil that is 1/10th the focal
length of the eyepiece, regardless of the dimensions of the objective.
Because magnification is proportional to focal length, but light grasp is proportional to the square of
aperture, the relative image illuminance produced by different relative apertures (different focal lengths
for the same aperture, or different apertures for the same focal length) scales as the ratio of the relative
apertures squared (diagram, left). Thus an /2 system has roughly six times the image illuminance of an
/5 system ((5/2)2 = 2.52 = 6.25), an /7 system has half the illuminance, an /10 system one fourth,
etc.
The standard system of photographic stops doubles the area of the diaphragm stop (effective aperture)
at each step, which means the stops increase the image illuminance by powers of the square root of
2 (1.414): 1, 1.414, 2, 2.83, 4, 5.65, 8, etc.

Exit Pupil
The exit window is the image
of the eyepiece field stop as
viewed through the eyepiece
(image, right). The exit pupil
is the image of the telescope
aperture stop viewed
through the eyepiece
(image, far right).
Although the exit pupil as
image is visible hovering
above or just inside the eye
lens of the eyepiece when viewed from an oblique angle, it is generally located at
the cross section of the telescope's afocal output where the projected, centered
aperture image appears physically smallest (as shown here). (This is called the
Ramsden disc in the United Kingdom, in honor of the 18th century optician who
first described it.) The exit pupil, as Ramsden disc, is the point where the
telescope afocal image is most concentrated and therefore brightest.
A commonly suggested procedure to measure the exit pupil always stated in
millimeters is with a reticule magnifier. This sometimes precedes the use of the
exit pupil in the calculation of system magnification as the ratio between objective
aperture (Do) and exit pupil diameter (de) that is, as the inverse of the beam
compression of the system:
[1] Mt = Do/de
However, measuring the size of the exit pupil is inconvenient, and it is simpler and
more intuitive to calculate magnification as the ratio of objective focal length (o)
to eyepiece focal length (e). Then the exit pupil diameter can be calculated as the
ratio between the objective aperture and the magnification produced by the
objective/eyepiece combination:
[2] de = Do/Mt
This seems to define the exit pupil is the ratio of aperture to magnification: an exit
pupil of 1.0mm means the aperture (in millimeters) is equal to the magnification;
an exit pupil of 2.5mm means the aperture is 2.5 times the magnification, and so
on. But that is a misinterpretation. In fact, this definition of the exit pupil is the
same proportion that defines the telescope image illuminance in relation to the
retinal illuminance of the naked eye:
It = (Do/Mt)2, therefore de = It / 2
which means the exit pupil is effectively
image illuminance standardized on pupil
aperture. And if the eye pupil is assumed
to be a personal constant then the exit
pupil is proportional to the perceived
image brightness.
The exit pupil is optically the smallest
aperture through which all the afocal light
from the telescope must pass, its diameter

is related to the luminance of the telescope image in the same way that the
opening in a camera diaphragm determines the amount of light that can enter a
camera (graph, left). However, the rising curve on the left side is accompanied by
an increasing true field of view and decreasing image magnification, which also
brighten the image. The "shutter" and magnification effects combine to brighten or
darken the image in inverse square proportion, and the difference in the
illuminance of the telescope image equals the ratio of two exit pupils squared:
[3] L = (de.2/de.1)2
The peak image brightness occurs when de = . At that point, the telescopic image
is exactly as bright as the naked eye image of the same area of sky (omitting
transmission losses in the instrument).
Because any larger exit pupil diameter will exceed the nominal eye pupil aperture:
de > . In this situation, the fixed eye pupil is the limiting factor. As the exit pupil
becomes larger, the image luminance declines because the area sampled by the
fixed pupil size is a smaller part of the total exit pupil area, effectively stopping
down the objective aperture.
However, if we substitute into formula [2] the alternative definition of
magnification (as o/e), we find a third definition of the exit pupil as the ratio
between the eyepiece focal length and the telescope relative aperture (No):
[4] de = Do/(o/e) = e/(o/Do) = e/No
An exit pupil of 1.0 mm is therefore produced by an eyepiece focal length that
equals the telescope relative aperture, a 3 mm exit pupil is produced by an
eyepiece focal length 3 times the telescope relative aperture, and so on. The exit
pupil is simply the eyepiece focal length projected by multiplication with a constant
value the relative aperture.
The value at which all elements form equalities is an exit pupil of 1.0:
de = 1.0, Mt = Do, e = No
where Do and e are in millimeters. Changing the exit pupil (which normally means
either changing the eyepiece e or stopping down the aperture Do) involves
reciprocal changes in the other two equalities, which leads to the general form of
exit pupil relationships:
[5] de = n, Mt = Do/n, e = nNo
The exit pupil by itself gives no information about the optical system, other than
the diameter of the Ramsden disc and the image brightness relative to the naked
eye brightness [as (de/)2]. The diagram (below) illustrates these exit pupil
calculations.

Three fixed physical quantities of the telescope system (objective focal length o,
eyepiece focal length e and objective aperture Do, shown in white, above) are
used to calcuate two system ratios (relative aperture No and magnification Mt,
both in orange). The fourth physical quantity, eye pupil , is assumed constant
and is only necessary to benchmark relative image brightness.
So long as the eyepiece focal length and system focal ratio remain constant, all
other physical parameters related to the telescope objective the aperture, focal
length, magnification, resolution limit and image scale can be changed without
changing the exit pupil.
Note specifically that the exit pupil has no fixed relationship to magnification: an
exit pupil of 2.0 mm can equal a magnification of 125x with a 250 mm (10")
aperture or a magnification of 40x with an 80 mm (3.2") aperture. As with the
eyepiece itself, the magnification produced by an exit pupil depends on the focal
length of the objective it is used with.
Since eyepieces are made to be swapped in an out of different optical systems, the
exit pupil is commonly used to choose an eyepiece magnification that is suitable
for the parameters of any specific telescope system it's easy math to multiply
focal ratio by the desired exit pupil value to get the necessary eyepiece focal
length. Handy in itself when using many different telescope systems, this
procedure is sometimes misapplied to the factitious twin "problems" of wasted
light (or wasted aperture) and empty magnification.
The astronomer prevents "wasted light" by using magnification that creates an exit
pupil that is not larger than the eye pupil. As explained above, the "problem" is
that an eye pupil smaller than the exit pupil will block some light from entering the
eye, effectively "stopping down" the objective aperture. But with most
commercially available astronomical equipment, it is difficult to exceed the eye
pupil aperture. With an /6 telescope and a pupil aperture of 6 mm (my measured
dark adapted pupil size), an eyepiece with focal length greater than 36 mm would
be necessary. Even in an /4 telescope, only eyepieces above 24mm would cause
the problem. And there is no "problem" if a longer focal length eyepiece is chosen
to give a wider, lower power view of an extended object. Then low magnification is
used to increase the true field of view at the "cost" of reduced image luminance, in
the same way that high magnification is used to increase visual resolution at the
"cost" of a smaller true field of view.
At the other extreme, "empty magnification" is said to be any image enlargement
greater than necessary to make the Airy disc just visible to the eye. Since the
linear width of the Airy disc is dependent only on the relative aperture (as
2.44No), it is always 0.00135No radians wide. This can be rewritten as the Airy
disc angular diameter [(0.00135 x 206265)No = 279No] as a ratio of the exit pupil:
s = 279/de arcseconds
Alternately, given your personal visual resolution limit of Rv (values cited in the
optical literature range from 70" to 280" or more), the corresponding exit pupil at
which the Airy disc will become visible is
[6] de = 279/Rv mm
which motivates the conclusion that an exit pupil de < 0.5 mm only provides
"empty magnification" because the Airy disc is then more than twice the size of the
most conservative resolution limit (280"). In fact, adopting the typical visual

resolution threshold (120") indicates diffraction artifacts are visible when the exit
pupil is as large as 2.3. But these abstract considerations ignore the complicating
factors of image motion (due to atmospheric turbulence), brightness and contrast,
all strongly affected by magnification. Exit pupils below 0.2 are routinely valuable
when resolving close, bright double stars, and William Herschel employed exit
pupils as small as 0.03 when studying double stars with his "most excellent" /13
reflector. There is no such thing as "empty magnification" only magnification
that is not optimal for the visual task at hand.
However the physical diameter of the Ramsden disc does have important visual
consequences. A small exit pupil (less than 1.0 mm) projects somewhat like a
pinhole, which can cause vivid shadowing of suspended internal debris or "floaters"
on the retina. At the same time the tiny Ramsden disc must be centered in the
pupil to produce a view of the entire eyepiece field, which becomes a kinesthetic
challenge in very wide angle eyepieces or awkward viewing postures unless the
observer has a chair or ladder for support. On the other hand, a smaller exit pupil
is only refracted by the central portion of the eye lens and cornea, minimizing the
effects of astigmatism. For these reasons alone, many older viewers prefer exit
pupils between 1-2 mm.
More significant problems with the exit pupil occur when observing during daylight
with a binocular or spotting telescope. Daylight images will contract the pupil
aperture to 2 mm or less, and this makes alignment of the eye pupil and exit
pupil(s) more difficult. It also makes any spherical aberration of the exit pupil
more noticeable, and minimizes aberrations produced by the observer's cornea
and lens.
Resolution
The resolution of an optical system is measured as the smallest angular width that
the system can image at the focal surface. In astronomical systems this limit is
determined by three things: the quality of the optics of the system, the turbulence
or distortion introduced by the atmosphere, and the quantum wave properties of
light. Only the last property can be calculated directly as the reciprocal of aperture,
1/Do.
The Diffraction Artifact. Analysis of the aperture resolution limit depends on a
quantum (physical) rather than geometrical analysis of light. If the light
wavefronts from a distant star are traveling in direction x parallel to the optical
axis of a telescope, then the width of each wavefront is measured along a
dimension y perpendicular to the direction of light. Because the star is so far away,
the scale of the y dimension is effectively infinite.
If we specify a specific energy of light by wavelength, such as the commonly used
= 555 nm (the average wavelength of a normal eye's peak photopic sensitivity),
then the quantum indeterminancy of the location and momentum of any photon in
the wavefront is distributed across the entire dimension y, and both the energy
and location of the photon can have a precise value, which means it can carry
information about the emitting source of the photons.
However, once the wavefront passes through the aperture stop of a telescope, all
the indeterminacy in the y dimension must be reduced to the aperture width D.
We require a fixed momentum and wavefront location for each photon at the focal
surface of the objective, because the photon contributes to the formation of an
image. But this fixed position must also include the indeterminacy contained in the
y dimension of the wavefront, which produces an "error" or uncertainty in the

specific location of the photon in the image. This quantum uncertainty is defined as
no smaller than:
U = h
where h is Planck's constant, 6.626 x 10-34, and is the frequency (energy) of the
light, and where the speed of light c is their product, c = . Casting these
quantum relationships in a form that defines the indeterminacy of a circular
aperture of diameter D yields:
UD = hc / hD = h / hD = / D
Thus, for a "green" wavelength of 555 nm in a 10" (254 mm) telescope, the
angular width of the indeterminacy becomes:
UD = 0.000555 / 254 = 2.185 x 10-6 radians
In the light from a far distant star, the result of the cumulative uncertainty from all
photons focused to a single image point does not have a simple "fuzzy" image
structure around the point, due to the destruction or reinforcement of light waves
as different parts of the wavefront are superimposed at the same location on the
image plane. Instead it appears as a central concentration of light surrounded by
one or more rings, the center and rings separated by concentric dark intervals.
The central concentration is called the Airy disc (after the Cambridge professor
George Biddell Airy who first described it mathematically in 1835), surrounded by
concentric diffraction rings. The angular dimension of these features, in relative
units of /D, the relative luminance (as a proportion of the peak Airy disc
luminance) and the percentage of total light energy enclosed by each dark ring are
given in the table (below).

dimension (k) in /D radians


feature

peak
percent of
illuminance total light

radius of
maximum

diameter of
maximum

Airy disc

1.0

first gap

1.22

2.44

83.8%

first ring

1.64

3.28

0.017

second gap

2.24

4.48

91.0%

second ring

2.66

5.32

0.0041

third gap

3.24

6.48

93.8%

third ring

3.90

7.80

0.0016

fourth gap

4.24

8.48

95.3%

the Airy disc and diffraction rings


of a "point" light source

Objective Resolution. The diffraction


artifact of Airy disc and encircling rings
is the smallest possible image of a
"point" light source. As a result, the
diffraction artifacts created by two
"incoherent" or physically unrelated
light sources (such as two stars)
separated by an arbitrarily small
angular distance will appear to merge,
because the angular separation
between them is smaller than the
angular diameter of the artifacts
produced by that aperture.
However, the spacing between two
diffraction artifacts, standardized on
the /Do radian width characteristic of
every aperture, can be scaled to
represent the minimum separation
between two point sources necessary
to clearly resolve or separate their
diffraction artifacts. This is the basis
for an angular resolution limit for point
light sources given the aperture Do.
In the Rayleigh resolution limit (diagram, right), the required spacing places the
peak intensity of each Airy disc over the first dark gap of the other star, which is
given as k = 1.22 or 1.22/D radians in the table above. This produces a merged
minimum brightness betwee the two stars that is roughly 70% of the peak
brightness of each Airy disc. The three most commonly used resolution limits in
astronomy are:
RR = 1.22/Do206265 arcseconds (Rayleigh criterion)
RD = 1.02/Do206265 arcseconds (Dawes criterion also approximately the Airy
disc full width half maximum or FWHM)
RS = 0.95/Do206265 arcseconds (Sparrow criterion)
where the factor 206265 is used to convert radians into arcseconds, Do is in
millimeters, and the wavelength of light is usually chosen to match the peak foveal
response, which is distributed over a rather broad range between approximately
530 to 580 nm.
Note that the more stringent Sparrow and Dawes resolution criteria also have a
more limited application. Both apply only to two stars of equal brightness viewed
through a circular telescope aperture with no central obstruction. In addition, the
Dawes resolution criterion is specified for matching 6th magnitude stars, and the
Sparrow criterion becomes larger as the magnitude difference between the two
stars increases. In contrast, the Rayleigh criterion is not affected by the absolute
or relative magnitude of the two stars and is not affected by any central
obstruction, and for those reasons is the criterion most generally used.
Adopting arcsecond measurement units and a reference wavelength of =
0.000555 mm, the more convenient formulas to calculate the resolution limits (in
aperture millimeters) are:

Ro = 140/Do (Rayleigh)
Ro = 117/Do (Dawes)
Ro = 109/Do (Sparrow)
and in aperture inches:
Ro = 5.50/Do (Rayleigh)
Ro = 4.60/Do (Dawes)
Ro = 4.28/Do (Sparrow).
The factors used in these formulas depend on the wavelength of light used as
reference. Thus, the Dawes criterion is more commonly given as 116/Do or
4.56/Do, which has been calculated on a wavelength of = 550 nm. These
differences are in practical application inconsequential.
In an afocal system with circular aperture, the unweighted quantum formula is
used to compute the minimum resolvable interval between two lines:
o = /Do206265 arcseconds
Thus, a diffraction limited 10" (254 mm) telescope can theoretically resolve two
lines separated by about 0.45 arcseconds, but two point sources (according to the
Rayleigh criterion) separated by about 0.55 arcseconds. "Theoretically" means
optical quality matters, and atmospheric turbulence intervenes.
Visual Resolution (With Eyepiece). The resolution delivered by the objective to the
image surface is only one part of the total optical system. Visual recognition of the
artifact, or any dimensional feature on the image surface, depends on its image
size relative to the resolution of the eyepiece/eye combination.
Unlike the angular width, the linear dimension of the diffraction artifact on the
image plane scales only with relative aperture, independent of aperture and focal
length:
o = kNo millimeters
where k is the dimensional of the feature radius or diameter taken from the table
above. For example, the linear diameter of the Airy disc (at = 0.000555 mm) in
an /8 telescope is 2.44 x 0.000555 x 8 = 0.011 mm.
This diffraction limited image dimension must be compared to the minimum
physical width that can be separated by the observer's eye with the assistance of a
specific eyepiece, defined as:
Re = eRv/206265 millimeters
where Rv is the visual resolution limit (in arcseconds, converted to radians by
division by 206265) and e is the eyepiece focal length (in millimeters). Thus the
minimum width that can be resolved with a 10 mm eyepiece, given a visual
resolution limit of Rv = 120 arcseconds, is 10 mm x 120"/206265 = 0.0058 mm.
Comparison of the two widths shows that the linear diameter of the Airy disk
(0.011 mm) is roughly twice the width of the resolution limit (0.0058 mm), and
therefore would be easily visible in an /8 optical system with a 10 mm eyepiece.
However this might not be the case if the eye were fully dark adapted and the

pupil aperture larger.


Appearance of the Diffraction Artifact.
Given a fixed aperture and focal ratio,
optical theory stipulates that
regardless of the magnitude of the
"point source" the diffraction artifact
will always have the same linear and
angular diameter and the same
proportional illuminances between the
Airy disc and any of the rings around it.
Remarkably, this does not correspond to the detailed visual appearance.
First, as Sidgwick (p.39) explains, "At normally encountered /ratios (say, /5 to
/20) no intensity gradient across the disc is perceptible, the border between the
disc and the first minimum appears nearly sharp, and the rings are brighter than
theory would indicate, the first ring being not much fainter than the disc itself. ...
The visible extent of the disc, like the number of rings visible, varies for a given
instrument with the brightness of the source, although the discs are in fact the
same size, irrespective of brightness."
The paradox of the varying disc diameter and number of rings was explained by
Airy as due to the luminance threshold of the eye. This is illustrated by plotting the
diffraction artifact using a log scale for the luminance (diagram, right), since the
relative brightness of lights viewed against a dark background scales as
b = b2/b1 = 100.4(m2-m1)
The dotted horizontal lines in the diagram show the relative effect of a visual
threshold that extends down to either ~1/10th to ~1/1000th the peak Airy disc
brightness. If a magnitude 2 star is bright enough to present three rings, this
explanation requires the third ring of a mag. 2 star to appear as bright as a
magnitude 9.5 star
(diagram below, A).
A threshold explanation for
the changing angular size
of the disc does not explain
the fact that the first
diffraction ring contracts
around this decreasing
disc, becoming slightly
thinner and separated from the disc by a constant or slightly narrower first gap
(diagram right, C). It does not keep the same diameter regardless of the star
magnitude, which would produce in very faint stars a visibly enlarged first gap
around a much reduced Airy disc (diagram right, B).
Magnification
Magnification is the increase in the angular size (apparent width) of an object in
comparison to the naked eye size of the object. In telescope optics, magnification
results from the combined effect of the telescope objective and eyepiece, with the
optics of the eye taken as the reference.
The eye has a variable focal length: it can focus objects at different distances
within the fixed distance from the cornea and lens to the retina. Which focal length

provides the reference? We use the distance that provides the maximum image
size on the retina, the minimum object distance at which an object will appear
comfortably in focus at full accommodation. For the average human observer this
near point is 250 mm (25 cm).

An object moved closer to the eye than the near point will appear larger, but also
out of focus. Focusing the eye on an object closer than 250 mm is made possible
by a magnifier lens or eyepiece.
The magnification is then the ratio between the angular size of the magnified
object and its angular size as viewed by the naked eye at the near point distance
of 250 mm. The distance size rule states that the ratio between the two distances
is the inverse of the ratio between the two visual angles. Thus, an object imaged
at 50 mm will appear magnified by 250/50 = 5 times, and will subtend a visual
angle tan() = 5 x tan().
The eyepiece effective focal length (EFL) is determined as the distance between
the eyepiece focal plane and the eyepiece principal plane, which is located where
the field angle of the exit window intercepts the radius of the field stop. This
distance must be divided into the visual angle of the field stop as it appears at
250 mm, so the eyepiece magnification (diagram below, right) is calculated as:
Me = 250/e
If the ocular is used as an actual magnifier, with the object held at the focal plane
of the eyepiece, then its magnification is 250/e+1.

The angle of the principal ray from an object through the telescope objective
remains constant, but this angle projects a greater field height at a longer focal
length. Therefore the telescope objective magnification is determined by its focal
length. Now, what telescope focal length will produce an image of a distant object
(such as the moon) that, when the image is viewed at the near point, will appear
exactly the same size as the naked eye image? The answer is: when o =
250 mm. Therefore the telescope magnification (diagram above, left) is equal to:
Mo = o/250

Because one magnification has a multiplying effect on the other, the magnification
of the combined telescope and eyepiece system is the product of the separate
telescope and eyepiece magnifications. The factor 250 cancels out, leaving the
ratio of the two focal lengths:
Mt = MoMe = o/250 x 250/e = o/e
Thus, in a 10" (254 mm) /8 telescope with a 20mm eyepiece, the magnification
is:
Mt = (254 x 8)/250 x 250/20 = 8.13 x 12.5 = 101.6
When caclulating magnification, keep in mind that manufacturer supplied data on
telescope and eyepiece focal lengths is often inexact, and therefore magnifications
calculated from them should be considered approximate. Magnifications greater
than about 10x (e < 25 mm) and free of serious aberrations are not possible with
a single lens; a compound lens is required.
Magnification affects the image illuminance: as magnification increases, a smaller
part of the sky (the true field of view) is passed to the eye, and therefore a
smaller quantity of light is being used to fill the apparent field of the eyepiece. In
other words, as magnification increases, the image becomes dimmer. These
effects can be calculated starting with the exit pupil equality de = Do/Mt = ,
assuming an eye pupil nominal diameter of 6 mm (table, below).

image
brightness

aperture (mm)
de
100 150 200

(naked eye = 1)

250

300

350

400

500

1.0

17

25

33

42

50

58

67

83

0.1

1.9

54

79

105

132

158

183

211

263

0.01

0.6 167 250

333

417

500

583

667

833

0.001

0.19 528 790 1053 1317 1580 1830 2107 2633

The telescopic image of an extended object (the Moon) can never be brighter than
the naked eye image of the same celestial area, due to "stopping down" of larger
beams by the eye pupil. The telescopic image of a star brightens with aperture,
because the larger quantity of light gathered by the larger aperture is concentrated
into a star image that does not increase in size (it remains a "point" source of
light).
Magnification Selection
Various guidelines have been proposed to determine the optimal or most useful
magnification to use with a telescope. These are generally stated as multiples of
the aperture (in millimeters or inches), as these determine the brightness (light
grasp) and detail (angular resolution) of the image to be examined: and
brighter, more detailed images respond better to high magnification.
The benefits of higher magnification are darkened sky background and better
visibility of faint stars, easier visual detection of image features at the resolution
limit, smaller field curvature and reduced optical aberrations. The drawbacks are a

smaller true field of view, decreased image luminance, amplified atmospheric


turbulence, exaggerated instrument vibrations, and greater difficulty centering and
manually tracking a target object. The pointlike projection of a tiny exit pupil can
also silhouette "floaters" in the eye and dust on the eyepiece.
The magnification necessary to visualize a given spatial frequency is the angular
resolution of the eye (Rv) divided by the resolution of the telescope system
(calculated above):
Mmin = Rv/Ro
where Rv is the minimum angular width that can be resolved by the naked eye,
quoted (depending on context) as ranging from 60" to 120" up to 360", depending
on the eye pupil diameter and the type of resolution required. This optimal
magnification range is usually supplemented with lower and higher powers, and
can be stated as simple multiples of aperture (for magnification) and reciprocal
multiples of relative aperture (for the corresponding eyepiece focal length). The
five common recommendations and the justifications offered for them are:
Mt = Do/, e = N Normal magnification This defines the magnification at
which the exit pupil equals the pupil aperture of the observer. The rationale is that
any lower magnification produces an exit pupil that is larger than the eye pupil,
which effectively stops down (reduces the light grasp) of the objective. But this is
a dubious rule. Few observers have actually measured their dark adapted pupil
aperture, and the customary value (8mm) is typically too large. (Mine is 5.9mm.)
The main objection is that low magnification has benefits unrelated to light grasp:
the true field of view is larger, pointing and tracking become easier and less
critical, and the effects of poor seeing are minimized. Finally, observing bright
objects such as the Moon can contract the pupil to 3mm or less and, with
commercially available eyepiece focal lengths (e 50mm), telescopes at relative
apertures above /6.3 cannot reach the prescribed "normal magnification"
anyway. A more practical guideline is:
Mt = 0.25Do to 0.33Do, e = 3N to 4N Minimum useful magnification The
lower bound at this magnification level is not determined by the diameter of the
eye pupil, but by the angular diameter of the target (star cluster, extended nebula,
etc.) and by the relative aperture of the telescope, which can introduce distracting
aberrations in wide field eyepieces at low powers. Note that magnification between
0.25Do and 0.5Do is useful for deep sky observing when a turbulent atmosphere
precludes detailed observation of planets or the Moon.
Mt = 0.5Do, e = 2.0N Minimum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the optimistic minimum angular resolution of the eye (i.e.,
~1 arcminute, or 60/Ro). Obviously, the quality of the observer's eyes affect this
rule significantly.
Mt = 1.0Do, e = 1.0N Maximum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the conservative minimum angular resolution of the eye (i.e., ~2
arcminutes, or 120/Ro). This is equivalent to "Whittaker's rule", usually stated as
Mt = 25Do when aperture is in inches (25 is just the number of millimeters in an
inch).
The "optimal magnification" rules are based on visual resolution limits, but it is

useful to observe with eyepieces in both focal lengths because an increase in


magnification causes sky brightness to decrease and very faint stars to become
more visible, despite the increase this may cause in the visible effects of
atmospheric turbulence.
Mt = 2.0Do to 4.0Do, e = 0.5N to 0.25N Maximum useful magnification
When the magnification causes the smallest (diffraction limited) image detail to be
just visible given the minimum angular resolution of the eye with a fully dark
adapted pupil (~5 arcminutes, or 300/Ro). This rule is highly dependent on
observing conditions and the observing target. Planetary and lunar observers will
generally prefer a lower magnification than double star observers, because
planetary details are low contrast and easily masked by poor seeing, whereas
stellar Airy discs tolerate high magnification very well, even under poor seeing.
In all cases, the maximum useful magnification has been exceeded when the
image is too dark (due to the magnification of the image), too blurry (due to the
quality of the optics, to atmospheric turbulence), or too difficult to view (due to
"floaters" in the observer's eye, small eye relief, etc.).
For the visual astronomer there is a less arbitrary approach to eyepiece selection.
Using a symmetrically cut aperture stop, shown here, a telescope trained on a
bright star will present a series of parallel bright and dark bands, spaced according
to the Rayleigh criterion (140/Do). View the bands with various eyepieces until you
find the eyepiece that yields a minimum but confident detection threshold. Divide
the telescope focal length by the eyepiece focal length to get the magnification,
then multiply the magnification by your telescope's Rayleigh resolution limit. This
yields your visual resolution threshold in arcseconds. (Mine is about 100.)
The simplest and most reliable rule is: use the magnification that produces the
most useful image. Atmospheric turbulence, the magnitude of stars or the surface
brightness and surface contrast of extended objects, the sky brightness, the type
and optical quality of the instrument, the quality of the observer's eyes and the
specific visual task or pastime all contribute to the choice of the optimal and most
useful magnification.
The corollary rule is: change the eyepiece often. No matter what the visual task or
observing target, trying different eyepieces will verify the best choice, and can
reveal unexpected details in the peripheral field or the target. Magnification
guidelines should be used merely guide the range of eyepieces selected for the
occasion.
Apparent Field of View
The apparent field of view (AFOV) of an optical system is the image of the field
stop in the image space, called the exit window. The AFOV is measured as the field
angle of the radius () or diameter (2) of the field edge or image limit (diagram,
below). The tangent of this angle is equal to the radius of the field stop opening
divided by the eyepiece back focal length (BFL).
The physical diameter of the eyepiece barrel and location of the field stop are
dictated by the optics of the eyepiece (its abaxial aberrations and design AFOV)
and the ratio of the objective (see here). The field stop excludes abaxial light
from the objective that would produce aberrations in the eyepiece image or an
unacceptably dark image, and blocks stray light reflected from the interior of the
telescope tube. The barrel diameter limits the maximum internal diameter of the
field stop, to ~27 mm in a 1.25" eyepiece barrel and ~48 mm in a standard 2"

eyepiece barrel. In all well made eyepieces, the field stop will define the apparent
field of view and will be crisply in focus.

The apparent field of view of an eyepiece is normally specified by the


manufacturer, along with the eyepiece focal length and design type. However,
the nominal AFOV of an eyepiece is often inflated due to the effects of positive
distortion, especially in "wide field" eyepieces (AFOV > 60).
When the manufacturer's apparent field specification appears inaccurate, or is not
known, then the AFOV can be measured in three ways:
(1) Projective Measurement. Support the eyepiece so that it is lying on its side
with the eye lens about 2 feet from a smooth wall or flat surface. (The seam
between pages of an open hardcover book, or between two books of the same
thickness laid spine to spine, is convenient.) Support a high powered flashlight in a
similar way so that it shines into the field lens (barrel end of the eyepiece) from a
distance of a few feet. (The eyepiece and flashlight should be lying perfectly level,
the optical axis of the eyepiece and flashlight must be perpendicular to the wall,
and the flashlight must be "collimated" or centered on and parallel to the eyepiece
optical axis, which can be arranged by centering the shadow of the flashlight bulb
inside the projected circle of light.) The eyepiece will project the light from the
flashlight onto the wall as a circle of light larger than the flashlight beam. Project
the image onto a small white card held close to the eye lens, and move the card
back and forth until the projected circle appears smallest: this is the location of the
exit pupil. Measure the greatest width of the circle (D) and the projection distance
(P) from the eyepiece exit pupil to the wall. Then:
AFOV = 2arctan(D/2P).
For example, if the imaged diameter is 82.3 cm, and the wall to exit pupil distance
is 73.3 cm, then:
AFOV = 2arctan(D/2P) = 2 x arctan(82.3/(2 x 73.3)) = 2 x 29.3 = 58.6.
(2) Visual Measurement. Tape to a convenient wall a long strip of paper or tape
measure on which small increments (inches or centimeters) are clearly marked
and visible from a distance (Dm) of about 100 cm. Hold the eyepiece over one eye
so that the full field is clearly visible and superimposed over the measurement
tape, visible with the other eye. Manipulate the eyepiece position until the tape
appears to measure the full diameter of the eyepiece field (De). Then:
AFOV = 2arctan(0.5De/Dm).

For example, viewing a metric measuring tape from 100 cm, you measure an
apparent width of the eyepiece field as 93 cm. Then:
AFOV = 2arctan(0.5 x 93/100) = 2arctan(0.47) = 2 x 25 = 50
(3) Physical Measurement. Use calipers to measure the interior diameter of the
field stop, assuming it exists and can be recognized. If in doubt, look through the
eyepiece and use the tip of a toothpick or pencil to locate the field stop: this is
either the end of the eyepiece barrel or the interior edge of the lock nut holding
the field lens. (Note that many "super wide field" eyepieces locate the field stop
inside the eyepiece where it cannot be measured.) Then the apparent field is
simply the visual angle that the field stop radius would subtend at 250 mm,
divided by the apparent field radius . The distance size equation then makes this
equal to:
AFOV = 2arctan(0.5Dfs/e).
For example, if the field stop diameter is 23mm and the eyepiece focal length is
20mm, then:
AFOV = 2arctan(0.5Dfs/e) = 2arctan(11.5/20) = 2arctan(0.58) = 2 x 29.9 =
59.8.
True Field of View
The exit window admits to the eye an image that has been magnified by both the
telescope and eyepiece. The actual size of the exit window, if the image area it
defines were viewed with the naked eye, would appear much smaller and much
farther away (diagram, above). This is the true field of view (TFOV) of the
objective/eyepiece combination.
Standard Formulas. The TFOV is equal to twice the maximum field radius admitted
to the image plane by the field stop:
TFOVo = 2max = 2h'max/o.
The TFOV is equal to the apparent field of view (AFOV) divided by the
magnification:
TFOV = AFOV/Mt.
Thus, if the apparent field of view of an eyepiece is 50, and the magnification of
the telescope and eyepiece combination is 200x, then the true field of view is:
TFOV = 50/200 = 0.25; or 0.25 x 60 = 15 arcminutes.
In the diagram (above), an eyepiece with a 50 field of view offers the same visual
area as a circular opening 23 cm in diameter held perpendicular to the eye at 25
cm distance. Then true field of view is equivalent to the area of sky seen through
the opening at a distance of 50 meters (164 feet); multiplication by 60 converts
degrees to arcminutes.
If the field stop diameter and objective focal length are known, then the TFOV can
be calculated as:

TFOV = Dfs/o x 57.3.


This formula also shows that the maximum true field of view possible with a 1.25"
eyepiece barrel or a visual back of a certain diameter. For a 10" /8 telescope, the
maximum true field of view is:
TFOV = 27/2032 x 57.3 = 0.76 x 60 = 46 arcminutes.
Star Drift Measurement. If the field stop diameter and apparent field of view are
not known, the true field of view can be measured by means of star transit times.
This requires at least three measurements of the time (in seconds) it takes a
bright star, placed outside the eyepiece field of view, to appear at one side of the
eyepiece field, cross the center of the field, and disappear at the opposite side.
Care must be taken to hold the head in a fixed position, so that the entire
circumference of the field edge is visible, and to pass the star through the center
of the eyepiece field.
Then the true field is the average of the three timings multiplied by the cosine of
the star's declination (stars closer to the celestial pole will require a longer time to
transit the eyepiece field):
TFOV = 0.25cos(Declstar) x t.
The factor 0.25 is necessary to convert from seconds in time to arcminutes of
angular width. For example, using Regulus (alpha Leonis) as the transit star
(Declination = +1158'), three eyepiece transit times are 134, 131 and 138
seconds. Then:
TFOV = 0.25cos(11.97)(134+131+138)/3 = 0.25 x 0.978 x 134.3 = 32.8
arcminutes.
Given the TFOV (in arcminutes) and system magnification, the AFOV (in degrees)
is:
AFOV = TFOV / 60Mt.
The TFOV determined by star transit measurements will typically not correspond
exactly to the TFOV calculated from the manufacturer supplied eyepiece apparent
field of view, and the magnification calculated from the manufacturer supplied
eyepiece focal length. This is usually because the eyepiece AFOV and e are
nominal, inferred from the computer optical design rather than the physical
construction of the eyepiece; and because timing star transits includes some
measurement error.
Effect of Central Obstruction
Resolution is affected by wave diffraction effects, but so too is the potential of the
objective to transmit contrast information at different spatial frequencies.
Transmission is specifically affected by a central obstruction, such as the
secondary mirror in a Newtonian or Cassegrainian design, which creates an
annular rather than circular aperture geometry and diffraction at four edges on the
diameter rather than just two. This obstruction ratio is defined as:
obstruction ratio () = ds / Do

where ds is the diameter of the secondary obstruction, the mounting for the
secondary mirror.
The standard method to evaluate optical performance across a range of spatial
frequencies is the modulation transfer function or MTF. The MTF shows the
proportion of luminance contrast that is retained in the image formed by an optical
system as the spatial frequency of the luminance variations is increased.
Geometrically, the MTF calculation is equivalent to the area of overlap between two
aperture superimposed at a displacement equal to the spatial frequency.

The stimulus used is either a pattern of crisp black and white lines (to evaluate the
transmission of sharp edges) or a sine wave pattern of modulated grays (to
evaluate the transmission of tonal gradients). Both methods converge on similar
conclusions.
Modulation transfer is defined as the reduction in image contrast between the
range of the luminance in the image (LI) as a proportion of the range of luminance
in the target or stimulus (LS):
contrast ratio = (LImax LImin) / (LSmax LSmin)
Because diffraction produces a tiny amount of smearing in the image, contrast is
imperceptibly reduced at low spatial frequencies (wide spacing) but becomes
severe at high spatial frequencies (very narrow spacing). So the MTF is scaled to
zoom in on the high spatial frequences, defined as the inverse of the angular
spacing in multiples of the minimum theoretically resolvable angular spacing
(/Dmm). For example, the Rayleigh criterion of resolution is defined as 1.22/D,
so its spatial frequency in an MTF is 1/1.22, or 0.82. (The MTF is sometimes
graphed in terms of line pairs per millimeter (lp/mm) for a specific telescope
relative aperture, but this produces the same focus on high frequency spacing and
the compression of low frequency spacing at the far left of the graph.)

The green line in the


generic MTF diagram
(right) shows the curve for
an optically perfect and
unobstructed circular
aperture transmitting a
sine wave pattern with a
spacing frequency up to
the resolution limit
which in a 250mm /10
telescope is a stimulus
angular separation of 0.45
arcseconds or a spacing on
the image plane of
0.0055 mm or 182 line
pairs per millimeter.
The yellow line shows the
effect on the contrast ratio
of a 25% obstruction,
typically the upper limit in
a Newtonian reflector, and
a 50% obstruction, the
upper limit in a Cassegrain derived design optimized for astrophotography (Ritchey
Chrtien or Schmidt Cassegrain). Both patterns can be summarized (reading the
graph left to right) as (1) a negligible (less than 10%) contrast reduction in spatial
frequencies from the full image diameter down to about 10 times the Rayleigh
limit, (2) a significant drop in contrast resolution down to about 2 or 3 times the
Rayleigh limit, (3) an increase in resolution that can exceed an unobstructed
aperture by a small amount (due to enhanced contrast caused by overlapping
wave reinforcement), and (4) resolution matching the theoretical minimum at the
Rayleigh limit.
Using the 50% obstruction curve as an example, the minimum resolution occurs at
around 3 times the minimum resolution, or an angular width of about 1.5
arcseconds. This means the most severe image degradation will occur in lunar
features that are about 3 kilometers wide, but will produce a slight crispening in
features half that size. It will most affect features that are about 3% of the width
of Jupiter's disk at opposition or about 6% the width of Mars's disk at opposition.
Contrast on the Galilean satellites of Jupiter and on the disks of Mercury, Uranus
and Neptune would be strongly obscured or obliterated. Double stars separated by
about 1.3 arcseconds would be most compromised, but very close pairs would be
easier to separate.
Obviously, whether and how much a central obstruction matters depends
somewhat on the size of the aperture, but more on the specific task the telescope
is used to perform. Astrophographers can be tolerant of a large central obstruction
because they typically work on an image scale of 1 pixel = 5x the resolution limit,
and image contrast can be enhanced with image stacking and digital manipulation.
For visual astronomers the problem is more complex. Inherently high contrast
targets such as the moon and double stars will display relatively minor contrast
reduction, and the effect will depend on magnification, because human contrast
sensitivity peaks at around 7 cycles per degree (around 10 arcminutes) and
declines as frequency increases. At the other extreme, low contrast and dim
targets such as deep sky objects will be minimally affected because the fovea

cannot function at their low luminances. This leaves planetary observation and
close double stars widely different in magnitude as the targets where the size of
the central obstruction can be a critical issue, one reason that Newtonian or
refractor telescopes are often preferred by planetary and double star observers.

Optics of the Eye


The eye is an equal component with objective and eyepiece in the performance of
an optical system. The critical performance dimensions for astronomy are dark
adaptation, hue discrimination, and resolution as a function of stimulus brightness
and stimulus/background contrast. However I begin with a topic that is customarily
minimized or ignored in discussions of visual performance: individual differences.
Individual Differences
Human vision is generally characterized in three ways: by describing the physical
structure and theoretical optics of the eye; by calculating the optical performance
of "ideal" or mathematical eyes; and by measuring individual performance in a
variety of visual tasks.
Visual performance is measured using tasks of detection (the stimulus is declared
to be visible or not visible), matching (two stimuli are declared to be or are
adjusted by the observer until they appear to be the same), or discrimination
(similar stimuli can be identified as different in some way). In all cases, statistical
theory is required to summarize performance across individuals. Stimulus
thresholds in detection or discrimination are defined as the stimulus value that
produces a 50% or 95% probability that an average person will respond with the
necessary detection or discrimination. Matching stimuli are defined by the average
and standard error of stimulus values judged to be the same, or the average just
noticeable difference (JND) that produces detection that they are different. These
measurements are then used to create models of average human performance,
which deviate from actual performance.
Two examples will illustrate the scale of the problems. On the subject of star color,
the first diagram (below left) shows the location on a standard hue circle of the
range (variation) in the single wavelengths chosen as a "pure" or best shade of
red, yellow, green and blue light by subjects with normal color vision; the inset
shows the number of individual subjects (circles) who chose a specific wavelength
in the "green" matching task. On the subject of telescope resolution limits, the
second diagram (below right) shows the variation in the judged brightness of
wavelengths across the visible spectrum; the "peak" wavelength is used to define
telescopic resolution limits.

In the color matching task, there seems to be good agreement among subjects in
the locus of a "pure" yellow wavelength in fact, this constitutes the Nagal
anomaloscope test for red/green color blindness. But in the green and blue color
matching the performance can best be described as scattered: the green values
chosen by different subjects range from a "green blue" below 500 nm to a "green
yellow" near 560 nm.
In terms of photopic luminance, individual
peak sensitivity spans a range of at least
40 nm (from 535 to 575 nm). The
commonly used value of 555 nm is the
peak of the 1964 CIE V* photopic
luminosity function (orange curve) or a
similar function, which are actually not
defined by tests of brightness matching
but by a weighted average of the L and M
cone response curves measured with color
mixture matching.
These individual differences appear in all
visual tasks, however they are measured. Peter Kaiser and Robert Boynton
describe the situation this way: "Comparisons between different observers,
whether in the same or a different experiment, present a discouraging picture.
Although observers agree on certain major trends, individual differences are best
described as enormous."
Standard psychometric methods are designed to disguise what is typically an
"enormous" amount of individual variation in perceptual thresholds of detection
and discrimination. These individual differences vary with the type of detection or
discrimination required, on the specific way in which the visual stimulus is created
and presented, and most importantly, on the perceptual and cognitive attributes of
the observer, including past experience with the task it is possible to learn or
practice your way to better visual acuity.
If you are curious about your visual performance as a visual astronomer, it will be
very helpful to discuss testing for color normal vision, resolution and contrast
sensitivity with your ophthalmologist. It is easy to measure pupil aperture
yourself, when your eye is fully dark adapted. This information will provide an
empirical basis for your assessment of telescopic optical performance.
Illuminance & Luminance
For the purposes of describing a light stimulus, the key distinctions are between
illuminance and luminance.
Illuminance (E) is the quantity of light incident on a unit area of surface or
passing through a unit area of aperture, defined as:
E = LI/d2
where LI is the luminous intensity of the light (the quantity of light radiated into
a solid angle of 1 steradian, in lumens) and d is the distance in from the light
source. When distance is in meters and the unit area is a 1 meter square surface
or 113 cm diameter circular aperture, illuminance is measured in lux.

Luminance is the quantity of light emitted by a surface onto a surface, defined as:
L = E/s
where s is the solid angle of the light source (in steradians or square radians) as
observed from the receiving surface or aperture. When both the light source
surface and the receptor surface are standardized as 1 meter square, luminance is
measured in candelas per square meter.
In optical systems, the objective focal length o and aperture area D2 and
reciprocal relative aperture govern the relationship between light grasp and image
luminance:
Lo = D2E/o2 = E/N2
As explained above, changes in relative aperture produce squared changes in
image luminance.
The key distinctions between illuminance and luminance are:
Illuminance is completely invisible to the eye; all light perception is a form of
imaged luminance, perceived either as an emitting light source or a reflecting
material. All brightness and lightness perceptions are luminance perceptions.
Illuminance decreases as the light source is at a larger distance from the
receiving surface. Luminance decreases as the same quantity of light is emitted by
a physically larger source. Illuminance and luminance both decrease as the same
amount of light is incident on a larger surface area.
Illuminance provides no information about the size, distance or intensity of the
light source (the same quantity of illuminance can come from a nearby dim light or
a distant bright light). Luminance is a specific measure of the visual geometry of
the illuminance from a light source in relation to a light receptor; it depends both
on the angular area (surface area and distance) of the light source and on the area
of the light receptor (e.g., the diameter of the pupil), but is invariant with distance
because a light source that decreases in illuminace with distance decreases in the
same inverse square proportion in angular area.
Illuminance decreases with increasing distance between the light source and
receiving surface. Luminance remains constant regardless of the distance between
the emitting light and receptor surface the angular area of the source decreases
with distance in the same inverse square proportion that the illuminance from the
source decreases.
The power or light concentrating capability of an optical system is proportional to
D2/2 = N2; a 2" aperture astrograph operating at /4 delivers more concentrated
light to a photoreceptor than an /10 instrument of 20" aperture. Regardless of
aperture, the luminance of an optical image is equal to the luminance of the
source.
Illuminance and luminance are measurement units in photometry, which is limited
to electromagnetic radiation that produces a response in the eye, usually
considered to be wavelengths between 750 to 380 nm. Irradiance and radiance are
the equivalent measurement units in radiometry, which includes all
electromagnetic energy from radio waves through xrays. Thus, a star that radiates

very powerfully but almost entirely in either the infrared or ultraviolet will appear
visually faint, because little of the energy from the star can stimulate the eye.
Luminance Adaptation
Vision can be characterized by three adaptation states, each associated with a
range of environmental light levels (illuminance):
Scotopic. Only the rods respond to light, generally at illuminance levels below 0.1
lux. Several visual criteria define this adaptation, chief among them is the loss of
color sensation, although memory color serves to tint familiar objects near the
scotopic threshold.
Photopic. The cones predominate in the response to light, which nominally occurs
at illuminance levels above 10 lux or more. In fact, the rods are still fully
functioning at illuminance levels even 10 times higher, up to light levels typical of
noon sunlight. Photopic adaptation characterizes the state in which color
perception and visual acuity can be adequately described in terms of the cones
alone.
Mesopic. The transition adaptation between scotopic and photopic vision. Color
perception is present, but color saturation and contrast are muted, and hues are
somewhat biased by the rods.
Light adaptation is the response of the eye to increased luminance levels. It is
generally prompt, occuring within a fraction of a second at the pupil and within a
minute in the retina and brain.
Dark adaptation is the response of the eye to decreased luminance levels. occurs
within 20 minutes to 1 hour, depending on the light levels.
Light Grasp
The light grasp of the eye as a function of aperture is quite modest. At a dark
adapted pupil diameter of about 6 mm, the pupil area allows only 3.5% the light
grasp of a 150 mm (6") telescope. However the eye is capable, under optimal
conditions, of perceiving a light source emitting only a few photons per second.
Thus, what it lacks in aperture it makes up in retinal sensitivity.
Resolution
The minimum angular interval that can be resolved by a normal human eye varies
with the type of resolution required, but is usually given as Rv = 70 to 140
arcseconds (0.00034 to 0.00068 radians). This is approximately 6 to 3 pairs of
alternating black/white lines within 1 millimeter viewed from 25cm (152 to 76 line
pairs per inch, viewed from 10 inches).
The next diagram describes the eye's optical resolution in the abstract, as the spot
diameter of a point light source on the retina as a function of pupil diameter,
based on calculations with the Navarro ideal or mathematical "eye".

The Navarro eye allows


the spot diameter to be
separated into the
contributions from the
size of the diffraction
artifact (the Airy disc,
which gets smaller with
larger pupil aperture)
and from optical
aberrations in the cornea
and lens (geometric
spreading, which
increases with aperture
larger than ~3 mm). At
the pupil sizes
characteristic of most
visual astronomy, the
spot diameter in a
normal eye is generally
far in excess of 25 m (0.025 mm); but note that a pupil diameter of around
3.5 mm would apply to most lunar and some planetary observing. Values will
generally be worse for older and visually impaired observers.
A common method for estimating the eye's peak visual acuity is to calculate the
average spacing between two cones in the fovea as the diameter of a single cone,
and to assume that visual resolution is equal to this average spacing (the gap
between two point stimuli is as wide as one cone). The average cone density is
estimated to be around 190,000 per mm2 in the foveola (diameter on the retina of
about 0.35 mm), and about 100,000 per mm2 in the fovea (diameter of about
1.85 mm). This works out to an average cone diameter of about 0.0025 mm in the
foveola and 0.0034 mm in the fovea, given the characteristic hexagonal "tiling" of
cone aperture.
With the eye's average internal focal length of 24mm, these widths define an
angular separation of between 0.006 to 0.0081, or 0.36 to 0.49 arcminutes.
However, the minimum spot diameter of the eye under optimal conditions
(diagram, above) is about 7 m or 0.007 mm, which is roughly the diameter of
three cones. Thus the cone spacing method for predicting visual acuity overstates
the eye's detection threshold under optimal conditions by a factor of roughly 3
and this is consistent with the measured performance acuity of the normal eye
under bright light, which is between 1 and 2 arcminutes across the fovea.

However, the spot


diameter under low
luminance adaptation is
well above 0.025 mm,
which implies a visual
acuity in most
observational astronomy
that is as poor as 10
arcminutes. This
degradation of visual
acuity with increasing
pupil aperture is clearly
indicated in the
modulation transfer
functions of normal eyes
at various pupil apertures,
measured by
photographing the spot
size actually imaged on
the retina of subjects
(diagram, below).
Finally, the task dependence of visual acuity is reflected in the different
discrimination thresholds that characterize different monocular visual tasks:
Shape recognition - Black on white letter shapes can be resolved (e.g.,
rotational orientations of the shape "E" can be identified) when elements of the
letter are at least 5 arcminutes wide.
Grating resolution - Black on white line gratings (the lines and the spaces
between them are of equal width) can be discriminated from a uniform gray
background when the spacing between black lines is at least 2 arcminutes.
Point recognition - Two black points on a white background can be resolved
when the space between them is at least 1 arcminute; white points on a black
background can be discriminated at about 2 arcminutes (due to spreading of the
image).
Vernier (Nonius) acuity - The misalignment between two black lines placed
end to end or the vertical misalignment of two dots can be perceived when the
lateral displacement is at least 10 arcseconds. This form of acuity is the principle
behind the heliometer and the reason why it was the tool that first measured
stellar parallax (Bessel and 61 Cygni).
Contrast Sensitivity
Contrast or modulation range

The troland is a measure of retinal illuminance equal to the illumination from a


surface emitting 1 lux per square meter into a pupil area of one square millimeter;
it increases both with an increase in source brightness and an increase in pupil
aperture. For comparison, 1000 trolands is approximately the brightness of a
"white" paper viewed under bright indoor lighting, and 0.1 trolands is
approximately the brightness of white paper viewed under moonlight.
Most "deep sky" astronomical targets present retinal illuminances of less than 0.1
trolands, which implies a visual acuity limited at around 2 to 5 cycles per degree or
Rv = 360 to 900 arcseconds.
Field of View
Because the eye's normal point is the standard for visual magnification, calculation
shows that the eye's true field of view is equal to its apparent field of view.
As a simple test of your true field of view: stand and look straight ahead with both
eyes, then extend your arms on either side, fingers facing forward. Wiggle your
fingers while moving both your arms forward and backward until you find the point
where the movement of the fingers of both hands is just visible on both sides of
your visual field. The angle between your arms will be about 120.
However, the fact that your fingers are barely visible at 120 means that your
useable field is much smaller. How small? At the minimum, the fovea is only able
to resolve the central ~1 of your visual field. Perceptual research also shows that
viewers will say a single object fills their visual field when it subtends about a 20
visual angle a generalization first made by Leonardo da Vinci. For example, the
industry recommended wide screen TV viewing distances produce a visual
angle to the wide screen diagonal of between 20 to 40, with the average around
25 an image width that is said to produce "an immersive feel". By that
standard, the 40 apparent field of traditional eyepieces would be fully immersive.
Currently available commercial eyepieces offer apparent fields of view between
40 to 100. The choice of one end of that range over the other seems to tip at an
apparent field of around 60, and can depend on several considerations: esthetic
preference (the "majesty factor" as opposed to "looking through a drinking
straw"), visual task (sweeping for comets as opposed to resolving double stars),
the telescope mounting and balance sensitivity (super wide field eyepieces can be
very heavy, especially at long eyepiece focal lengths; super wide field eyepieces
allow extended "drift" viewing of objects with hand guided Dobsonian altaz

mountings), objective relative aperture (off axis eyepiece aberrations are more
noticeable at smaller focal ratios) and price (super wide field eyepieces are
generally much more expensive than traditional designs).
Off axis image quality, in telescopes with focal ratios down to about /6, can be
superior in some super wide angle eyepieces than in some traditional eyepieces,
and image quality at the center of the field (on axis) is, for almost all oculars
operating at all focal ratios, optically perfect. Only at objective focal ratios below
/5 do optical considerations become important.
On the other hand, super wide field eyepieces require the observer to "look
around" in the wide available field in order to see clearly, and this movement of
the eye and head can cause vignetting or aberrations to appear. (Bias
disclosure: I prefer eyepieces with apparent fields no greater than 70, mainly to
avoid these effects.)
In general, preference for wide or "narrow" apparent field eyepieces is not an issue
of optical quality or the "immersive" visual area useful to the steady gaze of the
human eye, but depends on the astronomer's personal preferences in visual
esthetics, ergonomics and visual tasks, the convenience of use with a specific type
of telescope ... and price.
Focus Diopters
Diopters are used to measure the distance that a lens must be moved forward or
backward along the optical axis to produce a change in the image focus equal to
one diopter of the eye. This focus diopter is found as:
e2/1000 mm.
Thus the movement required to adjust the focus of a 20mm eyepiece by one
diopter is (202)/1000, or 0.4mm.
Focus diopters may be positive or negative. At zero focus diopter, an eyepiece
projects collimated light rays of light parallel to the optical axis, which can then
focused by a relaxed normal eye as if the light came from a far distant source.
Therefore the optimal focus diopter for a normal eye is zero or very slightly
negative.

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic
optical principles.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of
historically important eyepiece designs.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces,
and results from my collection.
Amateur Astronomer's Handbook by J.B. Sidgwick - excellent basic chapters in
optics, light gathering, resolution, magnification and more.

Telescopic Limiting Magnitudes by Bradley Schaefer - an attempt to predict


telescopic limiting magnitudes using visual data, star color, age of observer, sky
brightness and other factors.
Human Eye from Handbook of Optical Systems, Vol. 4: Survey of Optical
Instruments by Herbert Gross, Fritz Blechinger & Bertram Achtner (eds.). (Berlin,
DR: Wiley-VCH, 2008).
Visual Acuity - Summary of the various methods used to test the human visus.
Average optical performance of the human eye as a function of age in a
normal population by A Guirao, C Gonzlez, M Redondo, E Geraghty, S Norrby
and P Artal. Investigative Ophthalmology & Visual Science Jan. 1999, pp. 203-213.
The N.A.A. Telescope Calculator - handy and accurate calculator page for most
optical parameters of a telescope/eyepiece combination.
Eyepiece Focal Length Measurement - Jim Easterbrook explains how to
measure eyepiece focal lengths in grating projections.
The Collimation - an excellent optical discussion by Thierry Legault
What Is a MTF Curve? - Basics of the MTF.

Last revised 6/20/12 2012 Bruce MacEvoy

Exit Pupil
The exit window is the image
of the eyepiece field stop as
viewed through the eyepiece
(image, right). The exit pupil
is the image of the telescope
aperture stop viewed
through the eyepiece
(image, far right).
Although the exit pupil as
image is visible hovering
above or just inside the eye
lens of the eyepiece when viewed from an oblique angle, it is generally located at
the cross section of the telescope's afocal output where the projected, centered
aperture image appears physically smallest (as shown here). (This is called the
Ramsden disc in the United Kingdom, in honor of the 18th century optician who
first described it.) The exit pupil, as Ramsden disc, is the point where the
telescope afocal image is most concentrated and therefore brightest.
A commonly suggested procedure to measure the exit pupil always stated in
millimeters is with a reticule magnifier. This sometimes precedes the use of the
exit pupil in the calculation of system magnification as the ratio between objective
aperture (Do) and exit pupil diameter (de) that is, as the inverse of the beam
compression of the system:
[1] Mt = Do/de
However, measuring the size of the exit pupil is inconvenient, and it is simpler and
more intuitive to calculate magnification as the ratio of objective focal length (o)
to eyepiece focal length (e). Then the exit pupil diameter can be calculated as the
ratio between the objective aperture and the magnification produced by the
objective/eyepiece combination:
[2] de = Do/Mt
This seems to define the exit pupil is the ratio of aperture to magnification: an exit
pupil of 1.0mm means the aperture (in millimeters) is equal to the magnification;
an exit pupil of 2.5mm means the aperture is 2.5 times the magnification, and so
on. But that is a misinterpretation. In fact, this definition of the exit pupil is the
same proportion that defines the telescope image illuminance in relation to the
retinal illuminance of the naked eye:
It = (Do/Mt)2, therefore de = It / 2
which means the exit pupil is effectively
image illuminance standardized on pupil
aperture. And if the eye pupil is assumed
to be a personal constant then the exit
pupil is proportional to the perceived
image brightness.
The exit pupil is optically the smallest
aperture through which all the afocal light
from the telescope must pass, its diameter

is related to the luminance of the telescope image in the same way that the
opening in a camera diaphragm determines the amount of light that can enter a
camera (graph, left). However, the rising curve on the left side is accompanied by
an increasing true field of view and decreasing image magnification, which also
brighten the image. The "shutter" and magnification effects combine to brighten or
darken the image in inverse square proportion, and the difference in the
illuminance of the telescope image equals the ratio of two exit pupils squared:
[3] L = (de.2/de.1)2
The peak image brightness occurs when de = . At that point, the telescopic image
is exactly as bright as the naked eye image of the same area of sky (omitting
transmission losses in the instrument).
Because any larger exit pupil diameter will exceed the nominal eye pupil aperture:
de > . In this situation, the fixed eye pupil is the limiting factor. As the exit pupil
becomes larger, the image luminance declines because the area sampled by the
fixed pupil size is a smaller part of the total exit pupil area, effectively stopping
down the objective aperture.
However, if we substitute into formula [2] the alternative definition of
magnification (as o/e), we find a third definition of the exit pupil as the ratio
between the eyepiece focal length and the telescope relative aperture (No):
[4] de = Do/(o/e) = e/(o/Do) = e/No
An exit pupil of 1.0 mm is therefore produced by an eyepiece focal length that
equals the telescope relative aperture, a 3 mm exit pupil is produced by an
eyepiece focal length 3 times the telescope relative aperture, and so on. The exit
pupil is simply the eyepiece focal length projected by multiplication with a constant
value the relative aperture.
The value at which all elements form equalities is an exit pupil of 1.0:
de = 1.0, Mt = Do, e = No
where Do and e are in millimeters. Changing the exit pupil (which normally means
either changing the eyepiece e or stopping down the aperture Do) involves
reciprocal changes in the other two equalities, which leads to the general form of
exit pupil relationships:
[5] de = n, Mt = Do/n, e = nNo
The exit pupil by itself gives no information about the optical system, other than
the diameter of the Ramsden disc and the image brightness relative to the naked
eye brightness [as (de/)2]. The diagram (below) illustrates these exit pupil
calculations.

Three fixed physical quantities of the telescope system (objective focal length o,
eyepiece focal length e and objective aperture Do, shown in white, above) are
used to calcuate two system ratios (relative aperture No and magnification Mt,
both in orange). The fourth physical quantity, eye pupil , is assumed constant
and is only necessary to benchmark relative image brightness.
So long as the eyepiece focal length and system focal ratio remain constant, all
other physical parameters related to the telescope objective the aperture, focal
length, magnification, resolution limit and image scale can be changed without
changing the exit pupil.
Note specifically that the exit pupil has no fixed relationship to magnification: an
exit pupil of 2.0 mm can equal a magnification of 125x with a 250 mm (10")
aperture or a magnification of 40x with an 80 mm (3.2") aperture. As with the
eyepiece itself, the magnification produced by an exit pupil depends on the focal
length of the objective it is used with.
Since eyepieces are made to be swapped in an out of different optical systems, the
exit pupil is commonly used to choose an eyepiece magnification that is suitable
for the parameters of any specific telescope system it's easy math to multiply
focal ratio by the desired exit pupil value to get the necessary eyepiece focal
length. Handy in itself when using many different telescope systems, this
procedure is sometimes misapplied to the factitious twin "problems" of wasted
light (or wasted aperture) and empty magnification.
The astronomer prevents "wasted light" by using magnification that creates an exit
pupil that is not larger than the eye pupil. As explained above, the "problem" is
that an eye pupil smaller than the exit pupil will block some light from entering the
eye, effectively "stopping down" the objective aperture. But with most
commercially available astronomical equipment, it is difficult to exceed the eye
pupil aperture. With an /6 telescope and a pupil aperture of 6 mm (my measured
dark adapted pupil size), an eyepiece with focal length greater than 36 mm would
be necessary. Even in an /4 telescope, only eyepieces above 24mm would cause
the problem. And there is no "problem" if a longer focal length eyepiece is chosen
to give a wider, lower power view of an extended object. Then low magnification is
used to increase the true field of view at the "cost" of reduced image luminance, in
the same way that high magnification is used to increase visual resolution at the
"cost" of a smaller true field of view.
At the other extreme, "empty magnification" is said to be any image enlargement
greater than necessary to make the Airy disc just visible to the eye. Since the
linear width of the Airy disc is dependent only on the relative aperture (as
2.44No), it is always 0.00135No radians wide. This can be rewritten as the Airy
disc angular diameter [(0.00135 x 206265)No = 279No] as a ratio of the exit pupil:
s = 279/de arcseconds
Alternately, given your personal visual resolution limit of Rv (values cited in the
optical literature range from 70" to 280" or more), the corresponding exit pupil at
which the Airy disc will become visible is
[6] de = 279/Rv mm
which motivates the conclusion that an exit pupil de < 0.5 mm only provides
"empty magnification" because the Airy disc is then more than twice the size of the
most conservative resolution limit (280"). In fact, adopting the typical visual

resolution threshold (120") indicates diffraction artifacts are visible when the exit
pupil is as large as 2.3. But these abstract considerations ignore the complicating
factors of image motion (due to atmospheric turbulence), brightness and contrast,
all strongly affected by magnification. Exit pupils below 0.2 are routinely valuable
when resolving close, bright double stars, and William Herschel employed exit
pupils as small as 0.03 when studying double stars with his "most excellent" /13
reflector. There is no such thing as "empty magnification" only magnification
that is not optimal for the visual task at hand.
However the physical diameter of the Ramsden disc does have important visual
consequences. A small exit pupil (less than 1.0 mm) projects somewhat like a
pinhole, which can cause vivid shadowing of suspended internal debris or "floaters"
on the retina. At the same time the tiny Ramsden disc must be centered in the
pupil to produce a view of the entire eyepiece field, which becomes a kinesthetic
challenge in very wide angle eyepieces or awkward viewing postures unless the
observer has a chair or ladder for support. On the other hand, a smaller exit pupil
is only refracted by the central portion of the eye lens and cornea, minimizing the
effects of astigmatism. For these reasons alone, many older viewers prefer exit
pupils between 1-2 mm.
More significant problems with the exit pupil occur when observing during daylight
with a binocular or spotting telescope. Daylight images will contract the pupil
aperture to 2 mm or less, and this makes alignment of the eye pupil and exit
pupil(s) more difficult. It also makes any spherical aberration of the exit pupil
more noticeable, and minimizes aberrations produced by the observer's cornea
and lens.
Resolution
The resolution of an optical system is measured as the smallest angular width that
the system can image at the focal surface. In astronomical systems this limit is
determined by three things: the quality of the optics of the system, the turbulence
or distortion introduced by the atmosphere, and the quantum wave properties of
light. Only the last property can be calculated directly as the reciprocal of aperture,
1/Do.
The Diffraction Artifact. Analysis of the aperture resolution limit depends on a
quantum (physical) rather than geometrical analysis of light. If the light
wavefronts from a distant star are traveling in direction x parallel to the optical
axis of a telescope, then the width of each wavefront is measured along a
dimension y perpendicular to the direction of light. Because the star is so far away,
the scale of the y dimension is effectively infinite.
If we specify a specific energy of light by wavelength, such as the commonly used
= 555 nm (the average wavelength of a normal eye's peak photopic sensitivity),
then the quantum indeterminancy of the location and momentum of any photon in
the wavefront is distributed across the entire dimension y, and both the energy
and location of the photon can have a precise value, which means it can carry
information about the emitting source of the photons.
However, once the wavefront passes through the aperture stop of a telescope, all
the indeterminacy in the y dimension must be reduced to the aperture width D.
We require a fixed momentum and wavefront location for each photon at the focal
surface of the objective, because the photon contributes to the formation of an
image. But this fixed position must also include the indeterminacy contained in the
y dimension of the wavefront, which produces an "error" or uncertainty in the

specific location of the photon in the image. This quantum uncertainty is defined as
no smaller than:
U = h
where h is Planck's constant, 6.626 x 10-34, and is the frequency (energy) of the
light, and where the speed of light c is their product, c = . Casting these
quantum relationships in a form that defines the indeterminacy of a circular
aperture of diameter D yields:
UD = hc / hD = h / hD = / D
Thus, for a "green" wavelength of 555 nm in a 10" (254 mm) telescope, the
angular width of the indeterminacy becomes:
UD = 0.000555 / 254 = 2.185 x 10-6 radians
In the light from a far distant star, the result of the cumulative uncertainty from all
photons focused to a single image point does not have a simple "fuzzy" image
structure around the point, due to the destruction or reinforcement of light waves
as different parts of the wavefront are superimposed at the same location on the
image plane. Instead it appears as a central concentration of light surrounded by
one or more rings, the center and rings separated by concentric dark intervals.
The central concentration is called the Airy disc (after the Cambridge professor
George Biddell Airy who first described it mathematically in 1835), surrounded by
concentric diffraction rings. The angular dimension of these features, in relative
units of /D, the relative luminance (as a proportion of the peak Airy disc
luminance) and the percentage of total light energy enclosed by each dark ring are
given in the table (below).

dimension (k) in /D radians


feature

peak
percent of
illuminance total light

radius of
maximum

diameter of
maximum

Airy disc

1.0

first gap

1.22

2.44

83.8%

first ring

1.64

3.28

0.017

second gap

2.24

4.48

91.0%

second ring

2.66

5.32

0.0041

third gap

3.24

6.48

93.8%

third ring

3.90

7.80

0.0016

fourth gap

4.24

8.48

95.3%

the Airy disc and diffraction rings


of a "point" light source

Objective Resolution. The diffraction


artifact of Airy disc and encircling rings
is the smallest possible image of a
"point" light source. As a result, the
diffraction artifacts created by two
"incoherent" or physically unrelated
light sources (such as two stars)
separated by an arbitrarily small
angular distance will appear to merge,
because the angular separation
between them is smaller than the
angular diameter of the artifacts
produced by that aperture.
However, the spacing between two
diffraction artifacts, standardized on
the /Do radian width characteristic of
every aperture, can be scaled to
represent the minimum separation
between two point sources necessary
to clearly resolve or separate their
diffraction artifacts. This is the basis
for an angular resolution limit for point
light sources given the aperture Do.
In the Rayleigh resolution limit (diagram, right), the required spacing places the
peak intensity of each Airy disc over the first dark gap of the other star, which is
given as k = 1.22 or 1.22/D radians in the table above. This produces a merged
minimum brightness betwee the two stars that is roughly 70% of the peak
brightness of each Airy disc. The three most commonly used resolution limits in
astronomy are:
RR = 1.22/Do206265 arcseconds (Rayleigh criterion)
RD = 1.02/Do206265 arcseconds (Dawes criterion also approximately the Airy
disc full width half maximum or FWHM)
RS = 0.95/Do206265 arcseconds (Sparrow criterion)
where the factor 206265 is used to convert radians into arcseconds, Do is in
millimeters, and the wavelength of light is usually chosen to match the peak foveal
response, which is distributed over a rather broad range between approximately
530 to 580 nm.
Note that the more stringent Sparrow and Dawes resolution criteria also have a
more limited application. Both apply only to two stars of equal brightness viewed
through a circular telescope aperture with no central obstruction. In addition, the
Dawes resolution criterion is specified for matching 6th magnitude stars, and the
Sparrow criterion becomes larger as the magnitude difference between the two
stars increases. In contrast, the Rayleigh criterion is not affected by the absolute
or relative magnitude of the two stars and is not affected by any central
obstruction, and for those reasons is the criterion most generally used.
Adopting arcsecond measurement units and a reference wavelength of =
0.000555 mm, the more convenient formulas to calculate the resolution limits (in
aperture millimeters) are:

Ro = 140/Do (Rayleigh)
Ro = 117/Do (Dawes)
Ro = 109/Do (Sparrow)
and in aperture inches:
Ro = 5.50/Do (Rayleigh)
Ro = 4.60/Do (Dawes)
Ro = 4.28/Do (Sparrow).
The factors used in these formulas depend on the wavelength of light used as
reference. Thus, the Dawes criterion is more commonly given as 116/Do or
4.56/Do, which has been calculated on a wavelength of = 550 nm. These
differences are in practical application inconsequential.
In an afocal system with circular aperture, the unweighted quantum formula is
used to compute the minimum resolvable interval between two lines:
o = /Do206265 arcseconds
Thus, a diffraction limited 10" (254 mm) telescope can theoretically resolve two
lines separated by about 0.45 arcseconds, but two point sources (according to the
Rayleigh criterion) separated by about 0.55 arcseconds. "Theoretically" means
optical quality matters, and atmospheric turbulence intervenes.
Visual Resolution (With Eyepiece). The resolution delivered by the objective to the
image surface is only one part of the total optical system. Visual recognition of the
artifact, or any dimensional feature on the image surface, depends on its image
size relative to the resolution of the eyepiece/eye combination.
Unlike the angular width, the linear dimension of the diffraction artifact on the
image plane scales only with relative aperture, independent of aperture and focal
length:
o = kNo millimeters
where k is the dimensional of the feature radius or diameter taken from the table
above. For example, the linear diameter of the Airy disc (at = 0.000555 mm) in
an /8 telescope is 2.44 x 0.000555 x 8 = 0.011 mm.
This diffraction limited image dimension must be compared to the minimum
physical width that can be separated by the observer's eye with the assistance of a
specific eyepiece, defined as:
Re = eRv/206265 millimeters
where Rv is the visual resolution limit (in arcseconds, converted to radians by
division by 206265) and e is the eyepiece focal length (in millimeters). Thus the
minimum width that can be resolved with a 10 mm eyepiece, given a visual
resolution limit of Rv = 120 arcseconds, is 10 mm x 120"/206265 = 0.0058 mm.
Comparison of the two widths shows that the linear diameter of the Airy disk
(0.011 mm) is roughly twice the width of the resolution limit (0.0058 mm), and
therefore would be easily visible in an /8 optical system with a 10 mm eyepiece.
However this might not be the case if the eye were fully dark adapted and the

pupil aperture larger.


Appearance of the Diffraction Artifact.
Given a fixed aperture and focal ratio,
optical theory stipulates that
regardless of the magnitude of the
"point source" the diffraction artifact
will always have the same linear and
angular diameter and the same
proportional illuminances between the
Airy disc and any of the rings around it.
Remarkably, this does not correspond to the detailed visual appearance.
First, as Sidgwick (p.39) explains, "At normally encountered /ratios (say, /5 to
/20) no intensity gradient across the disc is perceptible, the border between the
disc and the first minimum appears nearly sharp, and the rings are brighter than
theory would indicate, the first ring being not much fainter than the disc itself. ...
The visible extent of the disc, like the number of rings visible, varies for a given
instrument with the brightness of the source, although the discs are in fact the
same size, irrespective of brightness."
The paradox of the varying disc diameter and number of rings was explained by
Airy as due to the luminance threshold of the eye. This is illustrated by plotting the
diffraction artifact using a log scale for the luminance (diagram, right), since the
relative brightness of lights viewed against a dark background scales as
b = b2/b1 = 100.4(m2-m1)
The dotted horizontal lines in the diagram show the relative effect of a visual
threshold that extends down to either ~1/10th to ~1/1000th the peak Airy disc
brightness. If a magnitude 2 star is bright enough to present three rings, this
explanation requires the third ring of a mag. 2 star to appear as bright as a
magnitude 9.5 star
(diagram below, A).
A threshold explanation for
the changing angular size
of the disc does not explain
the fact that the first
diffraction ring contracts
around this decreasing
disc, becoming slightly
thinner and separated from the disc by a constant or slightly narrower first gap
(diagram right, C). It does not keep the same diameter regardless of the star
magnitude, which would produce in very faint stars a visibly enlarged first gap
around a much reduced Airy disc (diagram right, B).
Magnification
Magnification is the increase in the angular size (apparent width) of an object in
comparison to the naked eye size of the object. In telescope optics, magnification
results from the combined effect of the telescope objective and eyepiece, with the
optics of the eye taken as the reference.
The eye has a variable focal length: it can focus objects at different distances
within the fixed distance from the cornea and lens to the retina. Which focal length

provides the reference? We use the distance that provides the maximum image
size on the retina, the minimum object distance at which an object will appear
comfortably in focus at full accommodation. For the average human observer this
near point is 250 mm (25 cm).

An object moved closer to the eye than the near point will appear larger, but also
out of focus. Focusing the eye on an object closer than 250 mm is made possible
by a magnifier lens or eyepiece.
The magnification is then the ratio between the angular size of the magnified
object and its angular size as viewed by the naked eye at the near point distance
of 250 mm. The distance size rule states that the ratio between the two distances
is the inverse of the ratio between the two visual angles. Thus, an object imaged
at 50 mm will appear magnified by 250/50 = 5 times, and will subtend a visual
angle tan() = 5 x tan().
The eyepiece effective focal length (EFL) is determined as the distance between
the eyepiece focal plane and the eyepiece principal plane, which is located where
the field angle of the exit window intercepts the radius of the field stop. This
distance must be divided into the visual angle of the field stop as it appears at
250 mm, so the eyepiece magnification (diagram below, right) is calculated as:
Me = 250/e
If the ocular is used as an actual magnifier, with the object held at the focal plane
of the eyepiece, then its magnification is 250/e+1.

The angle of the principal ray from an object through the telescope objective
remains constant, but this angle projects a greater field height at a longer focal
length. Therefore the telescope objective magnification is determined by its focal
length. Now, what telescope focal length will produce an image of a distant object
(such as the moon) that, when the image is viewed at the near point, will appear
exactly the same size as the naked eye image? The answer is: when o =
250 mm. Therefore the telescope magnification (diagram above, left) is equal to:
Mo = o/250

Because one magnification has a multiplying effect on the other, the magnification
of the combined telescope and eyepiece system is the product of the separate
telescope and eyepiece magnifications. The factor 250 cancels out, leaving the
ratio of the two focal lengths:
Mt = MoMe = o/250 x 250/e = o/e
Thus, in a 10" (254 mm) /8 telescope with a 20mm eyepiece, the magnification
is:
Mt = (254 x 8)/250 x 250/20 = 8.13 x 12.5 = 101.6
When caclulating magnification, keep in mind that manufacturer supplied data on
telescope and eyepiece focal lengths is often inexact, and therefore magnifications
calculated from them should be considered approximate. Magnifications greater
than about 10x (e < 25 mm) and free of serious aberrations are not possible with
a single lens; a compound lens is required.
Magnification affects the image illuminance: as magnification increases, a smaller
part of the sky (the true field of view) is passed to the eye, and therefore a
smaller quantity of light is being used to fill the apparent field of the eyepiece. In
other words, as magnification increases, the image becomes dimmer. These
effects can be calculated starting with the exit pupil equality de = Do/Mt = ,
assuming an eye pupil nominal diameter of 6 mm (table, below).

image
brightness

aperture (mm)
de
100 150 200

(naked eye = 1)

250

300

350

400

500

1.0

17

25

33

42

50

58

67

83

0.1

1.9

54

79

105

132

158

183

211

263

0.01

0.6 167 250

333

417

500

583

667

833

0.001

0.19 528 790 1053 1317 1580 1830 2107 2633

The telescopic image of an extended object (the Moon) can never be brighter than
the naked eye image of the same celestial area, due to "stopping down" of larger
beams by the eye pupil. The telescopic image of a star brightens with aperture,
because the larger quantity of light gathered by the larger aperture is concentrated
into a star image that does not increase in size (it remains a "point" source of
light).
Magnification Selection
Various guidelines have been proposed to determine the optimal or most useful
magnification to use with a telescope. These are generally stated as multiples of
the aperture (in millimeters or inches), as these determine the brightness (light
grasp) and detail (angular resolution) of the image to be examined: and
brighter, more detailed images respond better to high magnification.
The benefits of higher magnification are darkened sky background and better
visibility of faint stars, easier visual detection of image features at the resolution
limit, smaller field curvature and reduced optical aberrations. The drawbacks are a

smaller true field of view, decreased image luminance, amplified atmospheric


turbulence, exaggerated instrument vibrations, and greater difficulty centering and
manually tracking a target object. The pointlike projection of a tiny exit pupil can
also silhouette "floaters" in the eye and dust on the eyepiece.
The magnification necessary to visualize a given spatial frequency is the angular
resolution of the eye (Rv) divided by the resolution of the telescope system
(calculated above):
Mmin = Rv/Ro
where Rv is the minimum angular width that can be resolved by the naked eye,
quoted (depending on context) as ranging from 60" to 120" up to 360", depending
on the eye pupil diameter and the type of resolution required. This optimal
magnification range is usually supplemented with lower and higher powers, and
can be stated as simple multiples of aperture (for magnification) and reciprocal
multiples of relative aperture (for the corresponding eyepiece focal length). The
five common recommendations and the justifications offered for them are:
Mt = Do/, e = N Normal magnification This defines the magnification at
which the exit pupil equals the pupil aperture of the observer. The rationale is that
any lower magnification produces an exit pupil that is larger than the eye pupil,
which effectively stops down (reduces the light grasp) of the objective. But this is
a dubious rule. Few observers have actually measured their dark adapted pupil
aperture, and the customary value (8mm) is typically too large. (Mine is 5.9mm.)
The main objection is that low magnification has benefits unrelated to light grasp:
the true field of view is larger, pointing and tracking become easier and less
critical, and the effects of poor seeing are minimized. Finally, observing bright
objects such as the Moon can contract the pupil to 3mm or less and, with
commercially available eyepiece focal lengths (e 50mm), telescopes at relative
apertures above /6.3 cannot reach the prescribed "normal magnification"
anyway. A more practical guideline is:
Mt = 0.25Do to 0.33Do, e = 3N to 4N Minimum useful magnification The
lower bound at this magnification level is not determined by the diameter of the
eye pupil, but by the angular diameter of the target (star cluster, extended nebula,
etc.) and by the relative aperture of the telescope, which can introduce distracting
aberrations in wide field eyepieces at low powers. Note that magnification between
0.25Do and 0.5Do is useful for deep sky observing when a turbulent atmosphere
precludes detailed observation of planets or the Moon.
Mt = 0.5Do, e = 2.0N Minimum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the optimistic minimum angular resolution of the eye (i.e.,
~1 arcminute, or 60/Ro). Obviously, the quality of the observer's eyes affect this
rule significantly.
Mt = 1.0Do, e = 1.0N Maximum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the conservative minimum angular resolution of the eye (i.e., ~2
arcminutes, or 120/Ro). This is equivalent to "Whittaker's rule", usually stated as
Mt = 25Do when aperture is in inches (25 is just the number of millimeters in an
inch).
The "optimal magnification" rules are based on visual resolution limits, but it is

useful to observe with eyepieces in both focal lengths because an increase in


magnification causes sky brightness to decrease and very faint stars to become
more visible, despite the increase this may cause in the visible effects of
atmospheric turbulence.
Mt = 2.0Do to 4.0Do, e = 0.5N to 0.25N Maximum useful magnification
When the magnification causes the smallest (diffraction limited) image detail to be
just visible given the minimum angular resolution of the eye with a fully dark
adapted pupil (~5 arcminutes, or 300/Ro). This rule is highly dependent on
observing conditions and the observing target. Planetary and lunar observers will
generally prefer a lower magnification than double star observers, because
planetary details are low contrast and easily masked by poor seeing, whereas
stellar Airy discs tolerate high magnification very well, even under poor seeing.
In all cases, the maximum useful magnification has been exceeded when the
image is too dark (due to the magnification of the image), too blurry (due to the
quality of the optics, to atmospheric turbulence), or too difficult to view (due to
"floaters" in the observer's eye, small eye relief, etc.).
For the visual astronomer there is a less arbitrary approach to eyepiece selection.
Using a symmetrically cut aperture stop, shown here, a telescope trained on a
bright star will present a series of parallel bright and dark bands, spaced according
to the Rayleigh criterion (140/Do). View the bands with various eyepieces until you
find the eyepiece that yields a minimum but confident detection threshold. Divide
the telescope focal length by the eyepiece focal length to get the magnification,
then multiply the magnification by your telescope's Rayleigh resolution limit. This
yields your visual resolution threshold in arcseconds. (Mine is about 100.)
The simplest and most reliable rule is: use the magnification that produces the
most useful image. Atmospheric turbulence, the magnitude of stars or the surface
brightness and surface contrast of extended objects, the sky brightness, the type
and optical quality of the instrument, the quality of the observer's eyes and the
specific visual task or pastime all contribute to the choice of the optimal and most
useful magnification.
The corollary rule is: change the eyepiece often. No matter what the visual task or
observing target, trying different eyepieces will verify the best choice, and can
reveal unexpected details in the peripheral field or the target. Magnification
guidelines should be used merely guide the range of eyepieces selected for the
occasion.
Apparent Field of View
The apparent field of view (AFOV) of an optical system is the image of the field
stop in the image space, called the exit window. The AFOV is measured as the field
angle of the radius () or diameter (2) of the field edge or image limit (diagram,
below). The tangent of this angle is equal to the radius of the field stop opening
divided by the eyepiece back focal length (BFL).
The physical diameter of the eyepiece barrel and location of the field stop are
dictated by the optics of the eyepiece (its abaxial aberrations and design AFOV)
and the ratio of the objective (see here). The field stop excludes abaxial light
from the objective that would produce aberrations in the eyepiece image or an
unacceptably dark image, and blocks stray light reflected from the interior of the
telescope tube. The barrel diameter limits the maximum internal diameter of the
field stop, to ~27 mm in a 1.25" eyepiece barrel and ~48 mm in a standard 2"

eyepiece barrel. In all well made eyepieces, the field stop will define the apparent
field of view and will be crisply in focus.

The apparent field of view of an eyepiece is normally specified by the


manufacturer, along with the eyepiece focal length and design type. However,
the nominal AFOV of an eyepiece is often inflated due to the effects of positive
distortion, especially in "wide field" eyepieces (AFOV > 60).
When the manufacturer's apparent field specification appears inaccurate, or is not
known, then the AFOV can be measured in three ways:
(1) Projective Measurement. Support the eyepiece so that it is lying on its side
with the eye lens about 2 feet from a smooth wall or flat surface. (The seam
between pages of an open hardcover book, or between two books of the same
thickness laid spine to spine, is convenient.) Support a high powered flashlight in a
similar way so that it shines into the field lens (barrel end of the eyepiece) from a
distance of a few feet. (The eyepiece and flashlight should be lying perfectly level,
the optical axis of the eyepiece and flashlight must be perpendicular to the wall,
and the flashlight must be "collimated" or centered on and parallel to the eyepiece
optical axis, which can be arranged by centering the shadow of the flashlight bulb
inside the projected circle of light.) The eyepiece will project the light from the
flashlight onto the wall as a circle of light larger than the flashlight beam. Project
the image onto a small white card held close to the eye lens, and move the card
back and forth until the projected circle appears smallest: this is the location of the
exit pupil. Measure the greatest width of the circle (D) and the projection distance
(P) from the eyepiece exit pupil to the wall. Then:
AFOV = 2arctan(D/2P).
For example, if the imaged diameter is 82.3 cm, and the wall to exit pupil distance
is 73.3 cm, then:
AFOV = 2arctan(D/2P) = 2 x arctan(82.3/(2 x 73.3)) = 2 x 29.3 = 58.6.
(2) Visual Measurement. Tape to a convenient wall a long strip of paper or tape
measure on which small increments (inches or centimeters) are clearly marked
and visible from a distance (Dm) of about 100 cm. Hold the eyepiece over one eye
so that the full field is clearly visible and superimposed over the measurement
tape, visible with the other eye. Manipulate the eyepiece position until the tape
appears to measure the full diameter of the eyepiece field (De). Then:
AFOV = 2arctan(0.5De/Dm).

For example, viewing a metric measuring tape from 100 cm, you measure an
apparent width of the eyepiece field as 93 cm. Then:
AFOV = 2arctan(0.5 x 93/100) = 2arctan(0.47) = 2 x 25 = 50
(3) Physical Measurement. Use calipers to measure the interior diameter of the
field stop, assuming it exists and can be recognized. If in doubt, look through the
eyepiece and use the tip of a toothpick or pencil to locate the field stop: this is
either the end of the eyepiece barrel or the interior edge of the lock nut holding
the field lens. (Note that many "super wide field" eyepieces locate the field stop
inside the eyepiece where it cannot be measured.) Then the apparent field is
simply the visual angle that the field stop radius would subtend at 250 mm,
divided by the apparent field radius . The distance size equation then makes this
equal to:
AFOV = 2arctan(0.5Dfs/e).
For example, if the field stop diameter is 23mm and the eyepiece focal length is
20mm, then:
AFOV = 2arctan(0.5Dfs/e) = 2arctan(11.5/20) = 2arctan(0.58) = 2 x 29.9 =
59.8.
True Field of View
The exit window admits to the eye an image that has been magnified by both the
telescope and eyepiece. The actual size of the exit window, if the image area it
defines were viewed with the naked eye, would appear much smaller and much
farther away (diagram, above). This is the true field of view (TFOV) of the
objective/eyepiece combination.
Standard Formulas. The TFOV is equal to twice the maximum field radius admitted
to the image plane by the field stop:
TFOVo = 2max = 2h'max/o.
The TFOV is equal to the apparent field of view (AFOV) divided by the
magnification:
TFOV = AFOV/Mt.
Thus, if the apparent field of view of an eyepiece is 50, and the magnification of
the telescope and eyepiece combination is 200x, then the true field of view is:
TFOV = 50/200 = 0.25; or 0.25 x 60 = 15 arcminutes.
In the diagram (above), an eyepiece with a 50 field of view offers the same visual
area as a circular opening 23 cm in diameter held perpendicular to the eye at 25
cm distance. Then true field of view is equivalent to the area of sky seen through
the opening at a distance of 50 meters (164 feet); multiplication by 60 converts
degrees to arcminutes.
If the field stop diameter and objective focal length are known, then the TFOV can
be calculated as:

TFOV = Dfs/o x 57.3.


This formula also shows that the maximum true field of view possible with a 1.25"
eyepiece barrel or a visual back of a certain diameter. For a 10" /8 telescope, the
maximum true field of view is:
TFOV = 27/2032 x 57.3 = 0.76 x 60 = 46 arcminutes.
Star Drift Measurement. If the field stop diameter and apparent field of view are
not known, the true field of view can be measured by means of star transit times.
This requires at least three measurements of the time (in seconds) it takes a
bright star, placed outside the eyepiece field of view, to appear at one side of the
eyepiece field, cross the center of the field, and disappear at the opposite side.
Care must be taken to hold the head in a fixed position, so that the entire
circumference of the field edge is visible, and to pass the star through the center
of the eyepiece field.
Then the true field is the average of the three timings multiplied by the cosine of
the star's declination (stars closer to the celestial pole will require a longer time to
transit the eyepiece field):
TFOV = 0.25cos(Declstar) x t.
The factor 0.25 is necessary to convert from seconds in time to arcminutes of
angular width. For example, using Regulus (alpha Leonis) as the transit star
(Declination = +1158'), three eyepiece transit times are 134, 131 and 138
seconds. Then:
TFOV = 0.25cos(11.97)(134+131+138)/3 = 0.25 x 0.978 x 134.3 = 32.8
arcminutes.
Given the TFOV (in arcminutes) and system magnification, the AFOV (in degrees)
is:
AFOV = TFOV / 60Mt.
The TFOV determined by star transit measurements will typically not correspond
exactly to the TFOV calculated from the manufacturer supplied eyepiece apparent
field of view, and the magnification calculated from the manufacturer supplied
eyepiece focal length. This is usually because the eyepiece AFOV and e are
nominal, inferred from the computer optical design rather than the physical
construction of the eyepiece; and because timing star transits includes some
measurement error.
Effect of Central Obstruction
Resolution is affected by wave diffraction effects, but so too is the potential of the
objective to transmit contrast information at different spatial frequencies.
Transmission is specifically affected by a central obstruction, such as the
secondary mirror in a Newtonian or Cassegrainian design, which creates an
annular rather than circular aperture geometry and diffraction at four edges on the
diameter rather than just two. This obstruction ratio is defined as:
obstruction ratio () = ds / Do

where ds is the diameter of the secondary obstruction, the mounting for the
secondary mirror.
The standard method to evaluate optical performance across a range of spatial
frequencies is the modulation transfer function or MTF. The MTF shows the
proportion of luminance contrast that is retained in the image formed by an optical
system as the spatial frequency of the luminance variations is increased.
Geometrically, the MTF calculation is equivalent to the area of overlap between two
aperture superimposed at a displacement equal to the spatial frequency.

The stimulus used is either a pattern of crisp black and white lines (to evaluate the
transmission of sharp edges) or a sine wave pattern of modulated grays (to
evaluate the transmission of tonal gradients). Both methods converge on similar
conclusions.
Modulation transfer is defined as the reduction in image contrast between the
range of the luminance in the image (LI) as a proportion of the range of luminance
in the target or stimulus (LS):
contrast ratio = (LImax LImin) / (LSmax LSmin)
Because diffraction produces a tiny amount of smearing in the image, contrast is
imperceptibly reduced at low spatial frequencies (wide spacing) but becomes
severe at high spatial frequencies (very narrow spacing). So the MTF is scaled to
zoom in on the high spatial frequences, defined as the inverse of the angular
spacing in multiples of the minimum theoretically resolvable angular spacing
(/Dmm). For example, the Rayleigh criterion of resolution is defined as 1.22/D,
so its spatial frequency in an MTF is 1/1.22, or 0.82. (The MTF is sometimes
graphed in terms of line pairs per millimeter (lp/mm) for a specific telescope
relative aperture, but this produces the same focus on high frequency spacing and
the compression of low frequency spacing at the far left of the graph.)

The green line in the


generic MTF diagram
(right) shows the curve for
an optically perfect and
unobstructed circular
aperture transmitting a
sine wave pattern with a
spacing frequency up to
the resolution limit
which in a 250mm /10
telescope is a stimulus
angular separation of 0.45
arcseconds or a spacing on
the image plane of
0.0055 mm or 182 line
pairs per millimeter.
The yellow line shows the
effect on the contrast ratio
of a 25% obstruction,
typically the upper limit in
a Newtonian reflector, and
a 50% obstruction, the
upper limit in a Cassegrain derived design optimized for astrophotography (Ritchey
Chrtien or Schmidt Cassegrain). Both patterns can be summarized (reading the
graph left to right) as (1) a negligible (less than 10%) contrast reduction in spatial
frequencies from the full image diameter down to about 10 times the Rayleigh
limit, (2) a significant drop in contrast resolution down to about 2 or 3 times the
Rayleigh limit, (3) an increase in resolution that can exceed an unobstructed
aperture by a small amount (due to enhanced contrast caused by overlapping
wave reinforcement), and (4) resolution matching the theoretical minimum at the
Rayleigh limit.
Using the 50% obstruction curve as an example, the minimum resolution occurs at
around 3 times the minimum resolution, or an angular width of about 1.5
arcseconds. This means the most severe image degradation will occur in lunar
features that are about 3 kilometers wide, but will produce a slight crispening in
features half that size. It will most affect features that are about 3% of the width
of Jupiter's disk at opposition or about 6% the width of Mars's disk at opposition.
Contrast on the Galilean satellites of Jupiter and on the disks of Mercury, Uranus
and Neptune would be strongly obscured or obliterated. Double stars separated by
about 1.3 arcseconds would be most compromised, but very close pairs would be
easier to separate.
Obviously, whether and how much a central obstruction matters depends
somewhat on the size of the aperture, but more on the specific task the telescope
is used to perform. Astrophographers can be tolerant of a large central obstruction
because they typically work on an image scale of 1 pixel = 5x the resolution limit,
and image contrast can be enhanced with image stacking and digital manipulation.
For visual astronomers the problem is more complex. Inherently high contrast
targets such as the moon and double stars will display relatively minor contrast
reduction, and the effect will depend on magnification, because human contrast
sensitivity peaks at around 7 cycles per degree (around 10 arcminutes) and
declines as frequency increases. At the other extreme, low contrast and dim
targets such as deep sky objects will be minimally affected because the fovea

cannot function at their low luminances. This leaves planetary observation and
close double stars widely different in magnitude as the targets where the size of
the central obstruction can be a critical issue, one reason that Newtonian or
refractor telescopes are often preferred by planetary and double star observers.

Optics of the Eye


The eye is an equal component with objective and eyepiece in the performance of
an optical system. The critical performance dimensions for astronomy are dark
adaptation, hue discrimination, and resolution as a function of stimulus brightness
and stimulus/background contrast. However I begin with a topic that is customarily
minimized or ignored in discussions of visual performance: individual differences.
Individual Differences
Human vision is generally characterized in three ways: by describing the physical
structure and theoretical optics of the eye; by calculating the optical performance
of "ideal" or mathematical eyes; and by measuring individual performance in a
variety of visual tasks.
Visual performance is measured using tasks of detection (the stimulus is declared
to be visible or not visible), matching (two stimuli are declared to be or are
adjusted by the observer until they appear to be the same), or discrimination
(similar stimuli can be identified as different in some way). In all cases, statistical
theory is required to summarize performance across individuals. Stimulus
thresholds in detection or discrimination are defined as the stimulus value that
produces a 50% or 95% probability that an average person will respond with the
necessary detection or discrimination. Matching stimuli are defined by the average
and standard error of stimulus values judged to be the same, or the average just
noticeable difference (JND) that produces detection that they are different. These
measurements are then used to create models of average human performance,
which deviate from actual performance.
Two examples will illustrate the scale of the problems. On the subject of star color,
the first diagram (below left) shows the location on a standard hue circle of the
range (variation) in the single wavelengths chosen as a "pure" or best shade of
red, yellow, green and blue light by subjects with normal color vision; the inset
shows the number of individual subjects (circles) who chose a specific wavelength
in the "green" matching task. On the subject of telescope resolution limits, the
second diagram (below right) shows the variation in the judged brightness of
wavelengths across the visible spectrum; the "peak" wavelength is used to define
telescopic resolution limits.

In the color matching task, there seems to be good agreement among subjects in
the locus of a "pure" yellow wavelength in fact, this constitutes the Nagal
anomaloscope test for red/green color blindness. But in the green and blue color
matching the performance can best be described as scattered: the green values
chosen by different subjects range from a "green blue" below 500 nm to a "green
yellow" near 560 nm.
In terms of photopic luminance, individual
peak sensitivity spans a range of at least
40 nm (from 535 to 575 nm). The
commonly used value of 555 nm is the
peak of the 1964 CIE V* photopic
luminosity function (orange curve) or a
similar function, which are actually not
defined by tests of brightness matching
but by a weighted average of the L and M
cone response curves measured with color
mixture matching.
These individual differences appear in all
visual tasks, however they are measured. Peter Kaiser and Robert Boynton
describe the situation this way: "Comparisons between different observers,
whether in the same or a different experiment, present a discouraging picture.
Although observers agree on certain major trends, individual differences are best
described as enormous."
Standard psychometric methods are designed to disguise what is typically an
"enormous" amount of individual variation in perceptual thresholds of detection
and discrimination. These individual differences vary with the type of detection or
discrimination required, on the specific way in which the visual stimulus is created
and presented, and most importantly, on the perceptual and cognitive attributes of
the observer, including past experience with the task it is possible to learn or
practice your way to better visual acuity.
If you are curious about your visual performance as a visual astronomer, it will be
very helpful to discuss testing for color normal vision, resolution and contrast
sensitivity with your ophthalmologist. It is easy to measure pupil aperture
yourself, when your eye is fully dark adapted. This information will provide an
empirical basis for your assessment of telescopic optical performance.
Illuminance & Luminance
For the purposes of describing a light stimulus, the key distinctions are between
illuminance and luminance.
Illuminance (E) is the quantity of light incident on a unit area of surface or
passing through a unit area of aperture, defined as:
E = LI/d2
where LI is the luminous intensity of the light (the quantity of light radiated into
a solid angle of 1 steradian, in lumens) and d is the distance in from the light
source. When distance is in meters and the unit area is a 1 meter square surface
or 113 cm diameter circular aperture, illuminance is measured in lux.

Luminance is the quantity of light emitted by a surface onto a surface, defined as:
L = E/s
where s is the solid angle of the light source (in steradians or square radians) as
observed from the receiving surface or aperture. When both the light source
surface and the receptor surface are standardized as 1 meter square, luminance is
measured in candelas per square meter.
In optical systems, the objective focal length o and aperture area D2 and
reciprocal relative aperture govern the relationship between light grasp and image
luminance:
Lo = D2E/o2 = E/N2
As explained above, changes in relative aperture produce squared changes in
image luminance.
The key distinctions between illuminance and luminance are:
Illuminance is completely invisible to the eye; all light perception is a form of
imaged luminance, perceived either as an emitting light source or a reflecting
material. All brightness and lightness perceptions are luminance perceptions.
Illuminance decreases as the light source is at a larger distance from the
receiving surface. Luminance decreases as the same quantity of light is emitted by
a physically larger source. Illuminance and luminance both decrease as the same
amount of light is incident on a larger surface area.
Illuminance provides no information about the size, distance or intensity of the
light source (the same quantity of illuminance can come from a nearby dim light or
a distant bright light). Luminance is a specific measure of the visual geometry of
the illuminance from a light source in relation to a light receptor; it depends both
on the angular area (surface area and distance) of the light source and on the area
of the light receptor (e.g., the diameter of the pupil), but is invariant with distance
because a light source that decreases in illuminace with distance decreases in the
same inverse square proportion in angular area.
Illuminance decreases with increasing distance between the light source and
receiving surface. Luminance remains constant regardless of the distance between
the emitting light and receptor surface the angular area of the source decreases
with distance in the same inverse square proportion that the illuminance from the
source decreases.
The power or light concentrating capability of an optical system is proportional to
D2/2 = N2; a 2" aperture astrograph operating at /4 delivers more concentrated
light to a photoreceptor than an /10 instrument of 20" aperture. Regardless of
aperture, the luminance of an optical image is equal to the luminance of the
source.
Illuminance and luminance are measurement units in photometry, which is limited
to electromagnetic radiation that produces a response in the eye, usually
considered to be wavelengths between 750 to 380 nm. Irradiance and radiance are
the equivalent measurement units in radiometry, which includes all
electromagnetic energy from radio waves through xrays. Thus, a star that radiates

very powerfully but almost entirely in either the infrared or ultraviolet will appear
visually faint, because little of the energy from the star can stimulate the eye.
Luminance Adaptation
Vision can be characterized by three adaptation states, each associated with a
range of environmental light levels (illuminance):
Scotopic. Only the rods respond to light, generally at illuminance levels below 0.1
lux. Several visual criteria define this adaptation, chief among them is the loss of
color sensation, although memory color serves to tint familiar objects near the
scotopic threshold.
Photopic. The cones predominate in the response to light, which nominally occurs
at illuminance levels above 10 lux or more. In fact, the rods are still fully
functioning at illuminance levels even 10 times higher, up to light levels typical of
noon sunlight. Photopic adaptation characterizes the state in which color
perception and visual acuity can be adequately described in terms of the cones
alone.
Mesopic. The transition adaptation between scotopic and photopic vision. Color
perception is present, but color saturation and contrast are muted, and hues are
somewhat biased by the rods.
Light adaptation is the response of the eye to increased luminance levels. It is
generally prompt, occuring within a fraction of a second at the pupil and within a
minute in the retina and brain.
Dark adaptation is the response of the eye to decreased luminance levels. occurs
within 20 minutes to 1 hour, depending on the light levels.
Light Grasp
The light grasp of the eye as a function of aperture is quite modest. At a dark
adapted pupil diameter of about 6 mm, the pupil area allows only 3.5% the light
grasp of a 150 mm (6") telescope. However the eye is capable, under optimal
conditions, of perceiving a light source emitting only a few photons per second.
Thus, what it lacks in aperture it makes up in retinal sensitivity.
Resolution
The minimum angular interval that can be resolved by a normal human eye varies
with the type of resolution required, but is usually given as Rv = 70 to 140
arcseconds (0.00034 to 0.00068 radians). This is approximately 6 to 3 pairs of
alternating black/white lines within 1 millimeter viewed from 25cm (152 to 76 line
pairs per inch, viewed from 10 inches).
The next diagram describes the eye's optical resolution in the abstract, as the spot
diameter of a point light source on the retina as a function of pupil diameter,
based on calculations with the Navarro ideal or mathematical "eye".

The Navarro eye allows


the spot diameter to be
separated into the
contributions from the
size of the diffraction
artifact (the Airy disc,
which gets smaller with
larger pupil aperture)
and from optical
aberrations in the cornea
and lens (geometric
spreading, which
increases with aperture
larger than ~3 mm). At
the pupil sizes
characteristic of most
visual astronomy, the
spot diameter in a
normal eye is generally
far in excess of 25 m (0.025 mm); but note that a pupil diameter of around
3.5 mm would apply to most lunar and some planetary observing. Values will
generally be worse for older and visually impaired observers.
A common method for estimating the eye's peak visual acuity is to calculate the
average spacing between two cones in the fovea as the diameter of a single cone,
and to assume that visual resolution is equal to this average spacing (the gap
between two point stimuli is as wide as one cone). The average cone density is
estimated to be around 190,000 per mm2 in the foveola (diameter on the retina of
about 0.35 mm), and about 100,000 per mm2 in the fovea (diameter of about
1.85 mm). This works out to an average cone diameter of about 0.0025 mm in the
foveola and 0.0034 mm in the fovea, given the characteristic hexagonal "tiling" of
cone aperture.
With the eye's average internal focal length of 24mm, these widths define an
angular separation of between 0.006 to 0.0081, or 0.36 to 0.49 arcminutes.
However, the minimum spot diameter of the eye under optimal conditions
(diagram, above) is about 7 m or 0.007 mm, which is roughly the diameter of
three cones. Thus the cone spacing method for predicting visual acuity overstates
the eye's detection threshold under optimal conditions by a factor of roughly 3
and this is consistent with the measured performance acuity of the normal eye
under bright light, which is between 1 and 2 arcminutes across the fovea.

However, the spot


diameter under low
luminance adaptation is
well above 0.025 mm,
which implies a visual
acuity in most
observational astronomy
that is as poor as 10
arcminutes. This
degradation of visual
acuity with increasing
pupil aperture is clearly
indicated in the
modulation transfer
functions of normal eyes
at various pupil apertures,
measured by
photographing the spot
size actually imaged on
the retina of subjects
(diagram, below).
Finally, the task dependence of visual acuity is reflected in the different
discrimination thresholds that characterize different monocular visual tasks:
Shape recognition - Black on white letter shapes can be resolved (e.g.,
rotational orientations of the shape "E" can be identified) when elements of the
letter are at least 5 arcminutes wide.
Grating resolution - Black on white line gratings (the lines and the spaces
between them are of equal width) can be discriminated from a uniform gray
background when the spacing between black lines is at least 2 arcminutes.
Point recognition - Two black points on a white background can be resolved
when the space between them is at least 1 arcminute; white points on a black
background can be discriminated at about 2 arcminutes (due to spreading of the
image).
Vernier (Nonius) acuity - The misalignment between two black lines placed
end to end or the vertical misalignment of two dots can be perceived when the
lateral displacement is at least 10 arcseconds. This form of acuity is the principle
behind the heliometer and the reason why it was the tool that first measured
stellar parallax (Bessel and 61 Cygni).
Contrast Sensitivity
Contrast or modulation range

The troland is a measure of retinal illuminance equal to the illumination from a


surface emitting 1 lux per square meter into a pupil area of one square millimeter;
it increases both with an increase in source brightness and an increase in pupil
aperture. For comparison, 1000 trolands is approximately the brightness of a
"white" paper viewed under bright indoor lighting, and 0.1 trolands is
approximately the brightness of white paper viewed under moonlight.
Most "deep sky" astronomical targets present retinal illuminances of less than 0.1
trolands, which implies a visual acuity limited at around 2 to 5 cycles per degree or
Rv = 360 to 900 arcseconds.
Field of View
Because the eye's normal point is the standard for visual magnification, calculation
shows that the eye's true field of view is equal to its apparent field of view.
As a simple test of your true field of view: stand and look straight ahead with both
eyes, then extend your arms on either side, fingers facing forward. Wiggle your
fingers while moving both your arms forward and backward until you find the point
where the movement of the fingers of both hands is just visible on both sides of
your visual field. The angle between your arms will be about 120.
However, the fact that your fingers are barely visible at 120 means that your
useable field is much smaller. How small? At the minimum, the fovea is only able
to resolve the central ~1 of your visual field. Perceptual research also shows that
viewers will say a single object fills their visual field when it subtends about a 20
visual angle a generalization first made by Leonardo da Vinci. For example, the
industry recommended wide screen TV viewing distances produce a visual
angle to the wide screen diagonal of between 20 to 40, with the average around
25 an image width that is said to produce "an immersive feel". By that
standard, the 40 apparent field of traditional eyepieces would be fully immersive.
Currently available commercial eyepieces offer apparent fields of view between
40 to 100. The choice of one end of that range over the other seems to tip at an
apparent field of around 60, and can depend on several considerations: esthetic
preference (the "majesty factor" as opposed to "looking through a drinking
straw"), visual task (sweeping for comets as opposed to resolving double stars),
the telescope mounting and balance sensitivity (super wide field eyepieces can be
very heavy, especially at long eyepiece focal lengths; super wide field eyepieces
allow extended "drift" viewing of objects with hand guided Dobsonian altaz

mountings), objective relative aperture (off axis eyepiece aberrations are more
noticeable at smaller focal ratios) and price (super wide field eyepieces are
generally much more expensive than traditional designs).
Off axis image quality, in telescopes with focal ratios down to about /6, can be
superior in some super wide angle eyepieces than in some traditional eyepieces,
and image quality at the center of the field (on axis) is, for almost all oculars
operating at all focal ratios, optically perfect. Only at objective focal ratios below
/5 do optical considerations become important.
On the other hand, super wide field eyepieces require the observer to "look
around" in the wide available field in order to see clearly, and this movement of
the eye and head can cause vignetting or aberrations to appear. (Bias
disclosure: I prefer eyepieces with apparent fields no greater than 70, mainly to
avoid these effects.)
In general, preference for wide or "narrow" apparent field eyepieces is not an issue
of optical quality or the "immersive" visual area useful to the steady gaze of the
human eye, but depends on the astronomer's personal preferences in visual
esthetics, ergonomics and visual tasks, the convenience of use with a specific type
of telescope ... and price.
Focus Diopters
Diopters are used to measure the distance that a lens must be moved forward or
backward along the optical axis to produce a change in the image focus equal to
one diopter of the eye. This focus diopter is found as:
e2/1000 mm.
Thus the movement required to adjust the focus of a 20mm eyepiece by one
diopter is (202)/1000, or 0.4mm.
Focus diopters may be positive or negative. At zero focus diopter, an eyepiece
projects collimated light rays of light parallel to the optical axis, which can then
focused by a relaxed normal eye as if the light came from a far distant source.
Therefore the optimal focus diopter for a normal eye is zero or very slightly
negative.

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic
optical principles.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of
historically important eyepiece designs.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces,
and results from my collection.
Amateur Astronomer's Handbook by J.B. Sidgwick - excellent basic chapters in
optics, light gathering, resolution, magnification and more.

Telescopic Limiting Magnitudes by Bradley Schaefer - an attempt to predict


telescopic limiting magnitudes using visual data, star color, age of observer, sky
brightness and other factors.
Human Eye from Handbook of Optical Systems, Vol. 4: Survey of Optical
Instruments by Herbert Gross, Fritz Blechinger & Bertram Achtner (eds.). (Berlin,
DR: Wiley-VCH, 2008).
Visual Acuity - Summary of the various methods used to test the human visus.
Average optical performance of the human eye as a function of age in a
normal population by A Guirao, C Gonzlez, M Redondo, E Geraghty, S Norrby
and P Artal. Investigative Ophthalmology & Visual Science Jan. 1999, pp. 203-213.
The N.A.A. Telescope Calculator - handy and accurate calculator page for most
optical parameters of a telescope/eyepiece combination.
Eyepiece Focal Length Measurement - Jim Easterbrook explains how to
measure eyepiece focal lengths in grating projections.
The Collimation - an excellent optical discussion by Thierry Legault
What Is a MTF Curve? - Basics of the MTF.

Last revised 6/20/12 2012 Bruce MacEvoy

Exit Pupil
The exit window is the image
of the eyepiece field stop as
viewed through the eyepiece
(image, right). The exit pupil
is the image of the telescope
aperture stop viewed
through the eyepiece
(image, far right).
Although the exit pupil as
image is visible hovering
above or just inside the eye
lens of the eyepiece when viewed from an oblique angle, it is generally located at
the cross section of the telescope's afocal output where the projected, centered
aperture image appears physically smallest (as shown here). (This is called the
Ramsden disc in the United Kingdom, in honor of the 18th century optician who
first described it.) The exit pupil, as Ramsden disc, is the point where the
telescope afocal image is most concentrated and therefore brightest.
A commonly suggested procedure to measure the exit pupil always stated in
millimeters is with a reticule magnifier. This sometimes precedes the use of the
exit pupil in the calculation of system magnification as the ratio between objective
aperture (Do) and exit pupil diameter (de) that is, as the inverse of the beam
compression of the system:
[1] Mt = Do/de
However, measuring the size of the exit pupil is inconvenient, and it is simpler and
more intuitive to calculate magnification as the ratio of objective focal length (o)
to eyepiece focal length (e). Then the exit pupil diameter can be calculated as the
ratio between the objective aperture and the magnification produced by the
objective/eyepiece combination:
[2] de = Do/Mt
This seems to define the exit pupil is the ratio of aperture to magnification: an exit
pupil of 1.0mm means the aperture (in millimeters) is equal to the magnification;
an exit pupil of 2.5mm means the aperture is 2.5 times the magnification, and so
on. But that is a misinterpretation. In fact, this definition of the exit pupil is the
same proportion that defines the telescope image illuminance in relation to the
retinal illuminance of the naked eye:
It = (Do/Mt)2, therefore de = It / 2
which means the exit pupil is effectively
image illuminance standardized on pupil
aperture. And if the eye pupil is assumed
to be a personal constant then the exit
pupil is proportional to the perceived
image brightness.
The exit pupil is optically the smallest
aperture through which all the afocal light
from the telescope must pass, its diameter

is related to the luminance of the telescope image in the same way that the
opening in a camera diaphragm determines the amount of light that can enter a
camera (graph, left). However, the rising curve on the left side is accompanied by
an increasing true field of view and decreasing image magnification, which also
brighten the image. The "shutter" and magnification effects combine to brighten or
darken the image in inverse square proportion, and the difference in the
illuminance of the telescope image equals the ratio of two exit pupils squared:
[3] L = (de.2/de.1)2
The peak image brightness occurs when de = . At that point, the telescopic image
is exactly as bright as the naked eye image of the same area of sky (omitting
transmission losses in the instrument).
Because any larger exit pupil diameter will exceed the nominal eye pupil aperture:
de > . In this situation, the fixed eye pupil is the limiting factor. As the exit pupil
becomes larger, the image luminance declines because the area sampled by the
fixed pupil size is a smaller part of the total exit pupil area, effectively stopping
down the objective aperture.
However, if we substitute into formula [2] the alternative definition of
magnification (as o/e), we find a third definition of the exit pupil as the ratio
between the eyepiece focal length and the telescope relative aperture (No):
[4] de = Do/(o/e) = e/(o/Do) = e/No
An exit pupil of 1.0 mm is therefore produced by an eyepiece focal length that
equals the telescope relative aperture, a 3 mm exit pupil is produced by an
eyepiece focal length 3 times the telescope relative aperture, and so on. The exit
pupil is simply the eyepiece focal length projected by multiplication with a constant
value the relative aperture.
The value at which all elements form equalities is an exit pupil of 1.0:
de = 1.0, Mt = Do, e = No
where Do and e are in millimeters. Changing the exit pupil (which normally means
either changing the eyepiece e or stopping down the aperture Do) involves
reciprocal changes in the other two equalities, which leads to the general form of
exit pupil relationships:
[5] de = n, Mt = Do/n, e = nNo
The exit pupil by itself gives no information about the optical system, other than
the diameter of the Ramsden disc and the image brightness relative to the naked
eye brightness [as (de/)2]. The diagram (below) illustrates these exit pupil
calculations.

Three fixed physical quantities of the telescope system (objective focal length o,
eyepiece focal length e and objective aperture Do, shown in white, above) are
used to calcuate two system ratios (relative aperture No and magnification Mt,
both in orange). The fourth physical quantity, eye pupil , is assumed constant
and is only necessary to benchmark relative image brightness.
So long as the eyepiece focal length and system focal ratio remain constant, all
other physical parameters related to the telescope objective the aperture, focal
length, magnification, resolution limit and image scale can be changed without
changing the exit pupil.
Note specifically that the exit pupil has no fixed relationship to magnification: an
exit pupil of 2.0 mm can equal a magnification of 125x with a 250 mm (10")
aperture or a magnification of 40x with an 80 mm (3.2") aperture. As with the
eyepiece itself, the magnification produced by an exit pupil depends on the focal
length of the objective it is used with.
Since eyepieces are made to be swapped in an out of different optical systems, the
exit pupil is commonly used to choose an eyepiece magnification that is suitable
for the parameters of any specific telescope system it's easy math to multiply
focal ratio by the desired exit pupil value to get the necessary eyepiece focal
length. Handy in itself when using many different telescope systems, this
procedure is sometimes misapplied to the factitious twin "problems" of wasted
light (or wasted aperture) and empty magnification.
The astronomer prevents "wasted light" by using magnification that creates an exit
pupil that is not larger than the eye pupil. As explained above, the "problem" is
that an eye pupil smaller than the exit pupil will block some light from entering the
eye, effectively "stopping down" the objective aperture. But with most
commercially available astronomical equipment, it is difficult to exceed the eye
pupil aperture. With an /6 telescope and a pupil aperture of 6 mm (my measured
dark adapted pupil size), an eyepiece with focal length greater than 36 mm would
be necessary. Even in an /4 telescope, only eyepieces above 24mm would cause
the problem. And there is no "problem" if a longer focal length eyepiece is chosen
to give a wider, lower power view of an extended object. Then low magnification is
used to increase the true field of view at the "cost" of reduced image luminance, in
the same way that high magnification is used to increase visual resolution at the
"cost" of a smaller true field of view.
At the other extreme, "empty magnification" is said to be any image enlargement
greater than necessary to make the Airy disc just visible to the eye. Since the
linear width of the Airy disc is dependent only on the relative aperture (as
2.44No), it is always 0.00135No radians wide. This can be rewritten as the Airy
disc angular diameter [(0.00135 x 206265)No = 279No] as a ratio of the exit pupil:
s = 279/de arcseconds
Alternately, given your personal visual resolution limit of Rv (values cited in the
optical literature range from 70" to 280" or more), the corresponding exit pupil at
which the Airy disc will become visible is
[6] de = 279/Rv mm
which motivates the conclusion that an exit pupil de < 0.5 mm only provides
"empty magnification" because the Airy disc is then more than twice the size of the
most conservative resolution limit (280"). In fact, adopting the typical visual

resolution threshold (120") indicates diffraction artifacts are visible when the exit
pupil is as large as 2.3. But these abstract considerations ignore the complicating
factors of image motion (due to atmospheric turbulence), brightness and contrast,
all strongly affected by magnification. Exit pupils below 0.2 are routinely valuable
when resolving close, bright double stars, and William Herschel employed exit
pupils as small as 0.03 when studying double stars with his "most excellent" /13
reflector. There is no such thing as "empty magnification" only magnification
that is not optimal for the visual task at hand.
However the physical diameter of the Ramsden disc does have important visual
consequences. A small exit pupil (less than 1.0 mm) projects somewhat like a
pinhole, which can cause vivid shadowing of suspended internal debris or "floaters"
on the retina. At the same time the tiny Ramsden disc must be centered in the
pupil to produce a view of the entire eyepiece field, which becomes a kinesthetic
challenge in very wide angle eyepieces or awkward viewing postures unless the
observer has a chair or ladder for support. On the other hand, a smaller exit pupil
is only refracted by the central portion of the eye lens and cornea, minimizing the
effects of astigmatism. For these reasons alone, many older viewers prefer exit
pupils between 1-2 mm.
More significant problems with the exit pupil occur when observing during daylight
with a binocular or spotting telescope. Daylight images will contract the pupil
aperture to 2 mm or less, and this makes alignment of the eye pupil and exit
pupil(s) more difficult. It also makes any spherical aberration of the exit pupil
more noticeable, and minimizes aberrations produced by the observer's cornea
and lens.
Resolution
The resolution of an optical system is measured as the smallest angular width that
the system can image at the focal surface. In astronomical systems this limit is
determined by three things: the quality of the optics of the system, the turbulence
or distortion introduced by the atmosphere, and the quantum wave properties of
light. Only the last property can be calculated directly as the reciprocal of aperture,
1/Do.
The Diffraction Artifact. Analysis of the aperture resolution limit depends on a
quantum (physical) rather than geometrical analysis of light. If the light
wavefronts from a distant star are traveling in direction x parallel to the optical
axis of a telescope, then the width of each wavefront is measured along a
dimension y perpendicular to the direction of light. Because the star is so far away,
the scale of the y dimension is effectively infinite.
If we specify a specific energy of light by wavelength, such as the commonly used
= 555 nm (the average wavelength of a normal eye's peak photopic sensitivity),
then the quantum indeterminancy of the location and momentum of any photon in
the wavefront is distributed across the entire dimension y, and both the energy
and location of the photon can have a precise value, which means it can carry
information about the emitting source of the photons.
However, once the wavefront passes through the aperture stop of a telescope, all
the indeterminacy in the y dimension must be reduced to the aperture width D.
We require a fixed momentum and wavefront location for each photon at the focal
surface of the objective, because the photon contributes to the formation of an
image. But this fixed position must also include the indeterminacy contained in the
y dimension of the wavefront, which produces an "error" or uncertainty in the

specific location of the photon in the image. This quantum uncertainty is defined as
no smaller than:
U = h
where h is Planck's constant, 6.626 x 10-34, and is the frequency (energy) of the
light, and where the speed of light c is their product, c = . Casting these
quantum relationships in a form that defines the indeterminacy of a circular
aperture of diameter D yields:
UD = hc / hD = h / hD = / D
Thus, for a "green" wavelength of 555 nm in a 10" (254 mm) telescope, the
angular width of the indeterminacy becomes:
UD = 0.000555 / 254 = 2.185 x 10-6 radians
In the light from a far distant star, the result of the cumulative uncertainty from all
photons focused to a single image point does not have a simple "fuzzy" image
structure around the point, due to the destruction or reinforcement of light waves
as different parts of the wavefront are superimposed at the same location on the
image plane. Instead it appears as a central concentration of light surrounded by
one or more rings, the center and rings separated by concentric dark intervals.
The central concentration is called the Airy disc (after the Cambridge professor
George Biddell Airy who first described it mathematically in 1835), surrounded by
concentric diffraction rings. The angular dimension of these features, in relative
units of /D, the relative luminance (as a proportion of the peak Airy disc
luminance) and the percentage of total light energy enclosed by each dark ring are
given in the table (below).

dimension (k) in /D radians


feature

peak
percent of
illuminance total light

radius of
maximum

diameter of
maximum

Airy disc

1.0

first gap

1.22

2.44

83.8%

first ring

1.64

3.28

0.017

second gap

2.24

4.48

91.0%

second ring

2.66

5.32

0.0041

third gap

3.24

6.48

93.8%

third ring

3.90

7.80

0.0016

fourth gap

4.24

8.48

95.3%

the Airy disc and diffraction rings


of a "point" light source

Objective Resolution. The diffraction


artifact of Airy disc and encircling rings
is the smallest possible image of a
"point" light source. As a result, the
diffraction artifacts created by two
"incoherent" or physically unrelated
light sources (such as two stars)
separated by an arbitrarily small
angular distance will appear to merge,
because the angular separation
between them is smaller than the
angular diameter of the artifacts
produced by that aperture.
However, the spacing between two
diffraction artifacts, standardized on
the /Do radian width characteristic of
every aperture, can be scaled to
represent the minimum separation
between two point sources necessary
to clearly resolve or separate their
diffraction artifacts. This is the basis
for an angular resolution limit for point
light sources given the aperture Do.
In the Rayleigh resolution limit (diagram, right), the required spacing places the
peak intensity of each Airy disc over the first dark gap of the other star, which is
given as k = 1.22 or 1.22/D radians in the table above. This produces a merged
minimum brightness betwee the two stars that is roughly 70% of the peak
brightness of each Airy disc. The three most commonly used resolution limits in
astronomy are:
RR = 1.22/Do206265 arcseconds (Rayleigh criterion)
RD = 1.02/Do206265 arcseconds (Dawes criterion also approximately the Airy
disc full width half maximum or FWHM)
RS = 0.95/Do206265 arcseconds (Sparrow criterion)
where the factor 206265 is used to convert radians into arcseconds, Do is in
millimeters, and the wavelength of light is usually chosen to match the peak foveal
response, which is distributed over a rather broad range between approximately
530 to 580 nm.
Note that the more stringent Sparrow and Dawes resolution criteria also have a
more limited application. Both apply only to two stars of equal brightness viewed
through a circular telescope aperture with no central obstruction. In addition, the
Dawes resolution criterion is specified for matching 6th magnitude stars, and the
Sparrow criterion becomes larger as the magnitude difference between the two
stars increases. In contrast, the Rayleigh criterion is not affected by the absolute
or relative magnitude of the two stars and is not affected by any central
obstruction, and for those reasons is the criterion most generally used.
Adopting arcsecond measurement units and a reference wavelength of =
0.000555 mm, the more convenient formulas to calculate the resolution limits (in
aperture millimeters) are:

Ro = 140/Do (Rayleigh)
Ro = 117/Do (Dawes)
Ro = 109/Do (Sparrow)
and in aperture inches:
Ro = 5.50/Do (Rayleigh)
Ro = 4.60/Do (Dawes)
Ro = 4.28/Do (Sparrow).
The factors used in these formulas depend on the wavelength of light used as
reference. Thus, the Dawes criterion is more commonly given as 116/Do or
4.56/Do, which has been calculated on a wavelength of = 550 nm. These
differences are in practical application inconsequential.
In an afocal system with circular aperture, the unweighted quantum formula is
used to compute the minimum resolvable interval between two lines:
o = /Do206265 arcseconds
Thus, a diffraction limited 10" (254 mm) telescope can theoretically resolve two
lines separated by about 0.45 arcseconds, but two point sources (according to the
Rayleigh criterion) separated by about 0.55 arcseconds. "Theoretically" means
optical quality matters, and atmospheric turbulence intervenes.
Visual Resolution (With Eyepiece). The resolution delivered by the objective to the
image surface is only one part of the total optical system. Visual recognition of the
artifact, or any dimensional feature on the image surface, depends on its image
size relative to the resolution of the eyepiece/eye combination.
Unlike the angular width, the linear dimension of the diffraction artifact on the
image plane scales only with relative aperture, independent of aperture and focal
length:
o = kNo millimeters
where k is the dimensional of the feature radius or diameter taken from the table
above. For example, the linear diameter of the Airy disc (at = 0.000555 mm) in
an /8 telescope is 2.44 x 0.000555 x 8 = 0.011 mm.
This diffraction limited image dimension must be compared to the minimum
physical width that can be separated by the observer's eye with the assistance of a
specific eyepiece, defined as:
Re = eRv/206265 millimeters
where Rv is the visual resolution limit (in arcseconds, converted to radians by
division by 206265) and e is the eyepiece focal length (in millimeters). Thus the
minimum width that can be resolved with a 10 mm eyepiece, given a visual
resolution limit of Rv = 120 arcseconds, is 10 mm x 120"/206265 = 0.0058 mm.
Comparison of the two widths shows that the linear diameter of the Airy disk
(0.011 mm) is roughly twice the width of the resolution limit (0.0058 mm), and
therefore would be easily visible in an /8 optical system with a 10 mm eyepiece.
However this might not be the case if the eye were fully dark adapted and the

pupil aperture larger.


Appearance of the Diffraction Artifact.
Given a fixed aperture and focal ratio,
optical theory stipulates that
regardless of the magnitude of the
"point source" the diffraction artifact
will always have the same linear and
angular diameter and the same
proportional illuminances between the
Airy disc and any of the rings around it.
Remarkably, this does not correspond to the detailed visual appearance.
First, as Sidgwick (p.39) explains, "At normally encountered /ratios (say, /5 to
/20) no intensity gradient across the disc is perceptible, the border between the
disc and the first minimum appears nearly sharp, and the rings are brighter than
theory would indicate, the first ring being not much fainter than the disc itself. ...
The visible extent of the disc, like the number of rings visible, varies for a given
instrument with the brightness of the source, although the discs are in fact the
same size, irrespective of brightness."
The paradox of the varying disc diameter and number of rings was explained by
Airy as due to the luminance threshold of the eye. This is illustrated by plotting the
diffraction artifact using a log scale for the luminance (diagram, right), since the
relative brightness of lights viewed against a dark background scales as
b = b2/b1 = 100.4(m2-m1)
The dotted horizontal lines in the diagram show the relative effect of a visual
threshold that extends down to either ~1/10th to ~1/1000th the peak Airy disc
brightness. If a magnitude 2 star is bright enough to present three rings, this
explanation requires the third ring of a mag. 2 star to appear as bright as a
magnitude 9.5 star
(diagram below, A).
A threshold explanation for
the changing angular size
of the disc does not explain
the fact that the first
diffraction ring contracts
around this decreasing
disc, becoming slightly
thinner and separated from the disc by a constant or slightly narrower first gap
(diagram right, C). It does not keep the same diameter regardless of the star
magnitude, which would produce in very faint stars a visibly enlarged first gap
around a much reduced Airy disc (diagram right, B).
Magnification
Magnification is the increase in the angular size (apparent width) of an object in
comparison to the naked eye size of the object. In telescope optics, magnification
results from the combined effect of the telescope objective and eyepiece, with the
optics of the eye taken as the reference.
The eye has a variable focal length: it can focus objects at different distances
within the fixed distance from the cornea and lens to the retina. Which focal length

provides the reference? We use the distance that provides the maximum image
size on the retina, the minimum object distance at which an object will appear
comfortably in focus at full accommodation. For the average human observer this
near point is 250 mm (25 cm).

An object moved closer to the eye than the near point will appear larger, but also
out of focus. Focusing the eye on an object closer than 250 mm is made possible
by a magnifier lens or eyepiece.
The magnification is then the ratio between the angular size of the magnified
object and its angular size as viewed by the naked eye at the near point distance
of 250 mm. The distance size rule states that the ratio between the two distances
is the inverse of the ratio between the two visual angles. Thus, an object imaged
at 50 mm will appear magnified by 250/50 = 5 times, and will subtend a visual
angle tan() = 5 x tan().
The eyepiece effective focal length (EFL) is determined as the distance between
the eyepiece focal plane and the eyepiece principal plane, which is located where
the field angle of the exit window intercepts the radius of the field stop. This
distance must be divided into the visual angle of the field stop as it appears at
250 mm, so the eyepiece magnification (diagram below, right) is calculated as:
Me = 250/e
If the ocular is used as an actual magnifier, with the object held at the focal plane
of the eyepiece, then its magnification is 250/e+1.

The angle of the principal ray from an object through the telescope objective
remains constant, but this angle projects a greater field height at a longer focal
length. Therefore the telescope objective magnification is determined by its focal
length. Now, what telescope focal length will produce an image of a distant object
(such as the moon) that, when the image is viewed at the near point, will appear
exactly the same size as the naked eye image? The answer is: when o =
250 mm. Therefore the telescope magnification (diagram above, left) is equal to:
Mo = o/250

Because one magnification has a multiplying effect on the other, the magnification
of the combined telescope and eyepiece system is the product of the separate
telescope and eyepiece magnifications. The factor 250 cancels out, leaving the
ratio of the two focal lengths:
Mt = MoMe = o/250 x 250/e = o/e
Thus, in a 10" (254 mm) /8 telescope with a 20mm eyepiece, the magnification
is:
Mt = (254 x 8)/250 x 250/20 = 8.13 x 12.5 = 101.6
When caclulating magnification, keep in mind that manufacturer supplied data on
telescope and eyepiece focal lengths is often inexact, and therefore magnifications
calculated from them should be considered approximate. Magnifications greater
than about 10x (e < 25 mm) and free of serious aberrations are not possible with
a single lens; a compound lens is required.
Magnification affects the image illuminance: as magnification increases, a smaller
part of the sky (the true field of view) is passed to the eye, and therefore a
smaller quantity of light is being used to fill the apparent field of the eyepiece. In
other words, as magnification increases, the image becomes dimmer. These
effects can be calculated starting with the exit pupil equality de = Do/Mt = ,
assuming an eye pupil nominal diameter of 6 mm (table, below).

image
brightness

aperture (mm)
de
100 150 200

(naked eye = 1)

250

300

350

400

500

1.0

17

25

33

42

50

58

67

83

0.1

1.9

54

79

105

132

158

183

211

263

0.01

0.6 167 250

333

417

500

583

667

833

0.001

0.19 528 790 1053 1317 1580 1830 2107 2633

The telescopic image of an extended object (the Moon) can never be brighter than
the naked eye image of the same celestial area, due to "stopping down" of larger
beams by the eye pupil. The telescopic image of a star brightens with aperture,
because the larger quantity of light gathered by the larger aperture is concentrated
into a star image that does not increase in size (it remains a "point" source of
light).
Magnification Selection
Various guidelines have been proposed to determine the optimal or most useful
magnification to use with a telescope. These are generally stated as multiples of
the aperture (in millimeters or inches), as these determine the brightness (light
grasp) and detail (angular resolution) of the image to be examined: and
brighter, more detailed images respond better to high magnification.
The benefits of higher magnification are darkened sky background and better
visibility of faint stars, easier visual detection of image features at the resolution
limit, smaller field curvature and reduced optical aberrations. The drawbacks are a

smaller true field of view, decreased image luminance, amplified atmospheric


turbulence, exaggerated instrument vibrations, and greater difficulty centering and
manually tracking a target object. The pointlike projection of a tiny exit pupil can
also silhouette "floaters" in the eye and dust on the eyepiece.
The magnification necessary to visualize a given spatial frequency is the angular
resolution of the eye (Rv) divided by the resolution of the telescope system
(calculated above):
Mmin = Rv/Ro
where Rv is the minimum angular width that can be resolved by the naked eye,
quoted (depending on context) as ranging from 60" to 120" up to 360", depending
on the eye pupil diameter and the type of resolution required. This optimal
magnification range is usually supplemented with lower and higher powers, and
can be stated as simple multiples of aperture (for magnification) and reciprocal
multiples of relative aperture (for the corresponding eyepiece focal length). The
five common recommendations and the justifications offered for them are:
Mt = Do/, e = N Normal magnification This defines the magnification at
which the exit pupil equals the pupil aperture of the observer. The rationale is that
any lower magnification produces an exit pupil that is larger than the eye pupil,
which effectively stops down (reduces the light grasp) of the objective. But this is
a dubious rule. Few observers have actually measured their dark adapted pupil
aperture, and the customary value (8mm) is typically too large. (Mine is 5.9mm.)
The main objection is that low magnification has benefits unrelated to light grasp:
the true field of view is larger, pointing and tracking become easier and less
critical, and the effects of poor seeing are minimized. Finally, observing bright
objects such as the Moon can contract the pupil to 3mm or less and, with
commercially available eyepiece focal lengths (e 50mm), telescopes at relative
apertures above /6.3 cannot reach the prescribed "normal magnification"
anyway. A more practical guideline is:
Mt = 0.25Do to 0.33Do, e = 3N to 4N Minimum useful magnification The
lower bound at this magnification level is not determined by the diameter of the
eye pupil, but by the angular diameter of the target (star cluster, extended nebula,
etc.) and by the relative aperture of the telescope, which can introduce distracting
aberrations in wide field eyepieces at low powers. Note that magnification between
0.25Do and 0.5Do is useful for deep sky observing when a turbulent atmosphere
precludes detailed observation of planets or the Moon.
Mt = 0.5Do, e = 2.0N Minimum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the optimistic minimum angular resolution of the eye (i.e.,
~1 arcminute, or 60/Ro). Obviously, the quality of the observer's eyes affect this
rule significantly.
Mt = 1.0Do, e = 1.0N Maximum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the conservative minimum angular resolution of the eye (i.e., ~2
arcminutes, or 120/Ro). This is equivalent to "Whittaker's rule", usually stated as
Mt = 25Do when aperture is in inches (25 is just the number of millimeters in an
inch).
The "optimal magnification" rules are based on visual resolution limits, but it is

useful to observe with eyepieces in both focal lengths because an increase in


magnification causes sky brightness to decrease and very faint stars to become
more visible, despite the increase this may cause in the visible effects of
atmospheric turbulence.
Mt = 2.0Do to 4.0Do, e = 0.5N to 0.25N Maximum useful magnification
When the magnification causes the smallest (diffraction limited) image detail to be
just visible given the minimum angular resolution of the eye with a fully dark
adapted pupil (~5 arcminutes, or 300/Ro). This rule is highly dependent on
observing conditions and the observing target. Planetary and lunar observers will
generally prefer a lower magnification than double star observers, because
planetary details are low contrast and easily masked by poor seeing, whereas
stellar Airy discs tolerate high magnification very well, even under poor seeing.
In all cases, the maximum useful magnification has been exceeded when the
image is too dark (due to the magnification of the image), too blurry (due to the
quality of the optics, to atmospheric turbulence), or too difficult to view (due to
"floaters" in the observer's eye, small eye relief, etc.).
For the visual astronomer there is a less arbitrary approach to eyepiece selection.
Using a symmetrically cut aperture stop, shown here, a telescope trained on a
bright star will present a series of parallel bright and dark bands, spaced according
to the Rayleigh criterion (140/Do). View the bands with various eyepieces until you
find the eyepiece that yields a minimum but confident detection threshold. Divide
the telescope focal length by the eyepiece focal length to get the magnification,
then multiply the magnification by your telescope's Rayleigh resolution limit. This
yields your visual resolution threshold in arcseconds. (Mine is about 100.)
The simplest and most reliable rule is: use the magnification that produces the
most useful image. Atmospheric turbulence, the magnitude of stars or the surface
brightness and surface contrast of extended objects, the sky brightness, the type
and optical quality of the instrument, the quality of the observer's eyes and the
specific visual task or pastime all contribute to the choice of the optimal and most
useful magnification.
The corollary rule is: change the eyepiece often. No matter what the visual task or
observing target, trying different eyepieces will verify the best choice, and can
reveal unexpected details in the peripheral field or the target. Magnification
guidelines should be used merely guide the range of eyepieces selected for the
occasion.
Apparent Field of View
The apparent field of view (AFOV) of an optical system is the image of the field
stop in the image space, called the exit window. The AFOV is measured as the field
angle of the radius () or diameter (2) of the field edge or image limit (diagram,
below). The tangent of this angle is equal to the radius of the field stop opening
divided by the eyepiece back focal length (BFL).
The physical diameter of the eyepiece barrel and location of the field stop are
dictated by the optics of the eyepiece (its abaxial aberrations and design AFOV)
and the ratio of the objective (see here). The field stop excludes abaxial light
from the objective that would produce aberrations in the eyepiece image or an
unacceptably dark image, and blocks stray light reflected from the interior of the
telescope tube. The barrel diameter limits the maximum internal diameter of the
field stop, to ~27 mm in a 1.25" eyepiece barrel and ~48 mm in a standard 2"

eyepiece barrel. In all well made eyepieces, the field stop will define the apparent
field of view and will be crisply in focus.

The apparent field of view of an eyepiece is normally specified by the


manufacturer, along with the eyepiece focal length and design type. However,
the nominal AFOV of an eyepiece is often inflated due to the effects of positive
distortion, especially in "wide field" eyepieces (AFOV > 60).
When the manufacturer's apparent field specification appears inaccurate, or is not
known, then the AFOV can be measured in three ways:
(1) Projective Measurement. Support the eyepiece so that it is lying on its side
with the eye lens about 2 feet from a smooth wall or flat surface. (The seam
between pages of an open hardcover book, or between two books of the same
thickness laid spine to spine, is convenient.) Support a high powered flashlight in a
similar way so that it shines into the field lens (barrel end of the eyepiece) from a
distance of a few feet. (The eyepiece and flashlight should be lying perfectly level,
the optical axis of the eyepiece and flashlight must be perpendicular to the wall,
and the flashlight must be "collimated" or centered on and parallel to the eyepiece
optical axis, which can be arranged by centering the shadow of the flashlight bulb
inside the projected circle of light.) The eyepiece will project the light from the
flashlight onto the wall as a circle of light larger than the flashlight beam. Project
the image onto a small white card held close to the eye lens, and move the card
back and forth until the projected circle appears smallest: this is the location of the
exit pupil. Measure the greatest width of the circle (D) and the projection distance
(P) from the eyepiece exit pupil to the wall. Then:
AFOV = 2arctan(D/2P).
For example, if the imaged diameter is 82.3 cm, and the wall to exit pupil distance
is 73.3 cm, then:
AFOV = 2arctan(D/2P) = 2 x arctan(82.3/(2 x 73.3)) = 2 x 29.3 = 58.6.
(2) Visual Measurement. Tape to a convenient wall a long strip of paper or tape
measure on which small increments (inches or centimeters) are clearly marked
and visible from a distance (Dm) of about 100 cm. Hold the eyepiece over one eye
so that the full field is clearly visible and superimposed over the measurement
tape, visible with the other eye. Manipulate the eyepiece position until the tape
appears to measure the full diameter of the eyepiece field (De). Then:
AFOV = 2arctan(0.5De/Dm).

For example, viewing a metric measuring tape from 100 cm, you measure an
apparent width of the eyepiece field as 93 cm. Then:
AFOV = 2arctan(0.5 x 93/100) = 2arctan(0.47) = 2 x 25 = 50
(3) Physical Measurement. Use calipers to measure the interior diameter of the
field stop, assuming it exists and can be recognized. If in doubt, look through the
eyepiece and use the tip of a toothpick or pencil to locate the field stop: this is
either the end of the eyepiece barrel or the interior edge of the lock nut holding
the field lens. (Note that many "super wide field" eyepieces locate the field stop
inside the eyepiece where it cannot be measured.) Then the apparent field is
simply the visual angle that the field stop radius would subtend at 250 mm,
divided by the apparent field radius . The distance size equation then makes this
equal to:
AFOV = 2arctan(0.5Dfs/e).
For example, if the field stop diameter is 23mm and the eyepiece focal length is
20mm, then:
AFOV = 2arctan(0.5Dfs/e) = 2arctan(11.5/20) = 2arctan(0.58) = 2 x 29.9 =
59.8.
True Field of View
The exit window admits to the eye an image that has been magnified by both the
telescope and eyepiece. The actual size of the exit window, if the image area it
defines were viewed with the naked eye, would appear much smaller and much
farther away (diagram, above). This is the true field of view (TFOV) of the
objective/eyepiece combination.
Standard Formulas. The TFOV is equal to twice the maximum field radius admitted
to the image plane by the field stop:
TFOVo = 2max = 2h'max/o.
The TFOV is equal to the apparent field of view (AFOV) divided by the
magnification:
TFOV = AFOV/Mt.
Thus, if the apparent field of view of an eyepiece is 50, and the magnification of
the telescope and eyepiece combination is 200x, then the true field of view is:
TFOV = 50/200 = 0.25; or 0.25 x 60 = 15 arcminutes.
In the diagram (above), an eyepiece with a 50 field of view offers the same visual
area as a circular opening 23 cm in diameter held perpendicular to the eye at 25
cm distance. Then true field of view is equivalent to the area of sky seen through
the opening at a distance of 50 meters (164 feet); multiplication by 60 converts
degrees to arcminutes.
If the field stop diameter and objective focal length are known, then the TFOV can
be calculated as:

TFOV = Dfs/o x 57.3.


This formula also shows that the maximum true field of view possible with a 1.25"
eyepiece barrel or a visual back of a certain diameter. For a 10" /8 telescope, the
maximum true field of view is:
TFOV = 27/2032 x 57.3 = 0.76 x 60 = 46 arcminutes.
Star Drift Measurement. If the field stop diameter and apparent field of view are
not known, the true field of view can be measured by means of star transit times.
This requires at least three measurements of the time (in seconds) it takes a
bright star, placed outside the eyepiece field of view, to appear at one side of the
eyepiece field, cross the center of the field, and disappear at the opposite side.
Care must be taken to hold the head in a fixed position, so that the entire
circumference of the field edge is visible, and to pass the star through the center
of the eyepiece field.
Then the true field is the average of the three timings multiplied by the cosine of
the star's declination (stars closer to the celestial pole will require a longer time to
transit the eyepiece field):
TFOV = 0.25cos(Declstar) x t.
The factor 0.25 is necessary to convert from seconds in time to arcminutes of
angular width. For example, using Regulus (alpha Leonis) as the transit star
(Declination = +1158'), three eyepiece transit times are 134, 131 and 138
seconds. Then:
TFOV = 0.25cos(11.97)(134+131+138)/3 = 0.25 x 0.978 x 134.3 = 32.8
arcminutes.
Given the TFOV (in arcminutes) and system magnification, the AFOV (in degrees)
is:
AFOV = TFOV / 60Mt.
The TFOV determined by star transit measurements will typically not correspond
exactly to the TFOV calculated from the manufacturer supplied eyepiece apparent
field of view, and the magnification calculated from the manufacturer supplied
eyepiece focal length. This is usually because the eyepiece AFOV and e are
nominal, inferred from the computer optical design rather than the physical
construction of the eyepiece; and because timing star transits includes some
measurement error.
Effect of Central Obstruction
Resolution is affected by wave diffraction effects, but so too is the potential of the
objective to transmit contrast information at different spatial frequencies.
Transmission is specifically affected by a central obstruction, such as the
secondary mirror in a Newtonian or Cassegrainian design, which creates an
annular rather than circular aperture geometry and diffraction at four edges on the
diameter rather than just two. This obstruction ratio is defined as:
obstruction ratio () = ds / Do

where ds is the diameter of the secondary obstruction, the mounting for the
secondary mirror.
The standard method to evaluate optical performance across a range of spatial
frequencies is the modulation transfer function or MTF. The MTF shows the
proportion of luminance contrast that is retained in the image formed by an optical
system as the spatial frequency of the luminance variations is increased.
Geometrically, the MTF calculation is equivalent to the area of overlap between two
aperture superimposed at a displacement equal to the spatial frequency.

The stimulus used is either a pattern of crisp black and white lines (to evaluate the
transmission of sharp edges) or a sine wave pattern of modulated grays (to
evaluate the transmission of tonal gradients). Both methods converge on similar
conclusions.
Modulation transfer is defined as the reduction in image contrast between the
range of the luminance in the image (LI) as a proportion of the range of luminance
in the target or stimulus (LS):
contrast ratio = (LImax LImin) / (LSmax LSmin)
Because diffraction produces a tiny amount of smearing in the image, contrast is
imperceptibly reduced at low spatial frequencies (wide spacing) but becomes
severe at high spatial frequencies (very narrow spacing). So the MTF is scaled to
zoom in on the high spatial frequences, defined as the inverse of the angular
spacing in multiples of the minimum theoretically resolvable angular spacing
(/Dmm). For example, the Rayleigh criterion of resolution is defined as 1.22/D,
so its spatial frequency in an MTF is 1/1.22, or 0.82. (The MTF is sometimes
graphed in terms of line pairs per millimeter (lp/mm) for a specific telescope
relative aperture, but this produces the same focus on high frequency spacing and
the compression of low frequency spacing at the far left of the graph.)

The green line in the


generic MTF diagram
(right) shows the curve for
an optically perfect and
unobstructed circular
aperture transmitting a
sine wave pattern with a
spacing frequency up to
the resolution limit
which in a 250mm /10
telescope is a stimulus
angular separation of 0.45
arcseconds or a spacing on
the image plane of
0.0055 mm or 182 line
pairs per millimeter.
The yellow line shows the
effect on the contrast ratio
of a 25% obstruction,
typically the upper limit in
a Newtonian reflector, and
a 50% obstruction, the
upper limit in a Cassegrain derived design optimized for astrophotography (Ritchey
Chrtien or Schmidt Cassegrain). Both patterns can be summarized (reading the
graph left to right) as (1) a negligible (less than 10%) contrast reduction in spatial
frequencies from the full image diameter down to about 10 times the Rayleigh
limit, (2) a significant drop in contrast resolution down to about 2 or 3 times the
Rayleigh limit, (3) an increase in resolution that can exceed an unobstructed
aperture by a small amount (due to enhanced contrast caused by overlapping
wave reinforcement), and (4) resolution matching the theoretical minimum at the
Rayleigh limit.
Using the 50% obstruction curve as an example, the minimum resolution occurs at
around 3 times the minimum resolution, or an angular width of about 1.5
arcseconds. This means the most severe image degradation will occur in lunar
features that are about 3 kilometers wide, but will produce a slight crispening in
features half that size. It will most affect features that are about 3% of the width
of Jupiter's disk at opposition or about 6% the width of Mars's disk at opposition.
Contrast on the Galilean satellites of Jupiter and on the disks of Mercury, Uranus
and Neptune would be strongly obscured or obliterated. Double stars separated by
about 1.3 arcseconds would be most compromised, but very close pairs would be
easier to separate.
Obviously, whether and how much a central obstruction matters depends
somewhat on the size of the aperture, but more on the specific task the telescope
is used to perform. Astrophographers can be tolerant of a large central obstruction
because they typically work on an image scale of 1 pixel = 5x the resolution limit,
and image contrast can be enhanced with image stacking and digital manipulation.
For visual astronomers the problem is more complex. Inherently high contrast
targets such as the moon and double stars will display relatively minor contrast
reduction, and the effect will depend on magnification, because human contrast
sensitivity peaks at around 7 cycles per degree (around 10 arcminutes) and
declines as frequency increases. At the other extreme, low contrast and dim
targets such as deep sky objects will be minimally affected because the fovea

cannot function at their low luminances. This leaves planetary observation and
close double stars widely different in magnitude as the targets where the size of
the central obstruction can be a critical issue, one reason that Newtonian or
refractor telescopes are often preferred by planetary and double star observers.

Optics of the Eye


The eye is an equal component with objective and eyepiece in the performance of
an optical system. The critical performance dimensions for astronomy are dark
adaptation, hue discrimination, and resolution as a function of stimulus brightness
and stimulus/background contrast. However I begin with a topic that is customarily
minimized or ignored in discussions of visual performance: individual differences.
Individual Differences
Human vision is generally characterized in three ways: by describing the physical
structure and theoretical optics of the eye; by calculating the optical performance
of "ideal" or mathematical eyes; and by measuring individual performance in a
variety of visual tasks.
Visual performance is measured using tasks of detection (the stimulus is declared
to be visible or not visible), matching (two stimuli are declared to be or are
adjusted by the observer until they appear to be the same), or discrimination
(similar stimuli can be identified as different in some way). In all cases, statistical
theory is required to summarize performance across individuals. Stimulus
thresholds in detection or discrimination are defined as the stimulus value that
produces a 50% or 95% probability that an average person will respond with the
necessary detection or discrimination. Matching stimuli are defined by the average
and standard error of stimulus values judged to be the same, or the average just
noticeable difference (JND) that produces detection that they are different. These
measurements are then used to create models of average human performance,
which deviate from actual performance.
Two examples will illustrate the scale of the problems. On the subject of star color,
the first diagram (below left) shows the location on a standard hue circle of the
range (variation) in the single wavelengths chosen as a "pure" or best shade of
red, yellow, green and blue light by subjects with normal color vision; the inset
shows the number of individual subjects (circles) who chose a specific wavelength
in the "green" matching task. On the subject of telescope resolution limits, the
second diagram (below right) shows the variation in the judged brightness of
wavelengths across the visible spectrum; the "peak" wavelength is used to define
telescopic resolution limits.

In the color matching task, there seems to be good agreement among subjects in
the locus of a "pure" yellow wavelength in fact, this constitutes the Nagal
anomaloscope test for red/green color blindness. But in the green and blue color
matching the performance can best be described as scattered: the green values
chosen by different subjects range from a "green blue" below 500 nm to a "green
yellow" near 560 nm.
In terms of photopic luminance, individual
peak sensitivity spans a range of at least
40 nm (from 535 to 575 nm). The
commonly used value of 555 nm is the
peak of the 1964 CIE V* photopic
luminosity function (orange curve) or a
similar function, which are actually not
defined by tests of brightness matching
but by a weighted average of the L and M
cone response curves measured with color
mixture matching.
These individual differences appear in all
visual tasks, however they are measured. Peter Kaiser and Robert Boynton
describe the situation this way: "Comparisons between different observers,
whether in the same or a different experiment, present a discouraging picture.
Although observers agree on certain major trends, individual differences are best
described as enormous."
Standard psychometric methods are designed to disguise what is typically an
"enormous" amount of individual variation in perceptual thresholds of detection
and discrimination. These individual differences vary with the type of detection or
discrimination required, on the specific way in which the visual stimulus is created
and presented, and most importantly, on the perceptual and cognitive attributes of
the observer, including past experience with the task it is possible to learn or
practice your way to better visual acuity.
If you are curious about your visual performance as a visual astronomer, it will be
very helpful to discuss testing for color normal vision, resolution and contrast
sensitivity with your ophthalmologist. It is easy to measure pupil aperture
yourself, when your eye is fully dark adapted. This information will provide an
empirical basis for your assessment of telescopic optical performance.
Illuminance & Luminance
For the purposes of describing a light stimulus, the key distinctions are between
illuminance and luminance.
Illuminance (E) is the quantity of light incident on a unit area of surface or
passing through a unit area of aperture, defined as:
E = LI/d2
where LI is the luminous intensity of the light (the quantity of light radiated into
a solid angle of 1 steradian, in lumens) and d is the distance in from the light
source. When distance is in meters and the unit area is a 1 meter square surface
or 113 cm diameter circular aperture, illuminance is measured in lux.

Luminance is the quantity of light emitted by a surface onto a surface, defined as:
L = E/s
where s is the solid angle of the light source (in steradians or square radians) as
observed from the receiving surface or aperture. When both the light source
surface and the receptor surface are standardized as 1 meter square, luminance is
measured in candelas per square meter.
In optical systems, the objective focal length o and aperture area D2 and
reciprocal relative aperture govern the relationship between light grasp and image
luminance:
Lo = D2E/o2 = E/N2
As explained above, changes in relative aperture produce squared changes in
image luminance.
The key distinctions between illuminance and luminance are:
Illuminance is completely invisible to the eye; all light perception is a form of
imaged luminance, perceived either as an emitting light source or a reflecting
material. All brightness and lightness perceptions are luminance perceptions.
Illuminance decreases as the light source is at a larger distance from the
receiving surface. Luminance decreases as the same quantity of light is emitted by
a physically larger source. Illuminance and luminance both decrease as the same
amount of light is incident on a larger surface area.
Illuminance provides no information about the size, distance or intensity of the
light source (the same quantity of illuminance can come from a nearby dim light or
a distant bright light). Luminance is a specific measure of the visual geometry of
the illuminance from a light source in relation to a light receptor; it depends both
on the angular area (surface area and distance) of the light source and on the area
of the light receptor (e.g., the diameter of the pupil), but is invariant with distance
because a light source that decreases in illuminace with distance decreases in the
same inverse square proportion in angular area.
Illuminance decreases with increasing distance between the light source and
receiving surface. Luminance remains constant regardless of the distance between
the emitting light and receptor surface the angular area of the source decreases
with distance in the same inverse square proportion that the illuminance from the
source decreases.
The power or light concentrating capability of an optical system is proportional to
D2/2 = N2; a 2" aperture astrograph operating at /4 delivers more concentrated
light to a photoreceptor than an /10 instrument of 20" aperture. Regardless of
aperture, the luminance of an optical image is equal to the luminance of the
source.
Illuminance and luminance are measurement units in photometry, which is limited
to electromagnetic radiation that produces a response in the eye, usually
considered to be wavelengths between 750 to 380 nm. Irradiance and radiance are
the equivalent measurement units in radiometry, which includes all
electromagnetic energy from radio waves through xrays. Thus, a star that radiates

very powerfully but almost entirely in either the infrared or ultraviolet will appear
visually faint, because little of the energy from the star can stimulate the eye.
Luminance Adaptation
Vision can be characterized by three adaptation states, each associated with a
range of environmental light levels (illuminance):
Scotopic. Only the rods respond to light, generally at illuminance levels below 0.1
lux. Several visual criteria define this adaptation, chief among them is the loss of
color sensation, although memory color serves to tint familiar objects near the
scotopic threshold.
Photopic. The cones predominate in the response to light, which nominally occurs
at illuminance levels above 10 lux or more. In fact, the rods are still fully
functioning at illuminance levels even 10 times higher, up to light levels typical of
noon sunlight. Photopic adaptation characterizes the state in which color
perception and visual acuity can be adequately described in terms of the cones
alone.
Mesopic. The transition adaptation between scotopic and photopic vision. Color
perception is present, but color saturation and contrast are muted, and hues are
somewhat biased by the rods.
Light adaptation is the response of the eye to increased luminance levels. It is
generally prompt, occuring within a fraction of a second at the pupil and within a
minute in the retina and brain.
Dark adaptation is the response of the eye to decreased luminance levels. occurs
within 20 minutes to 1 hour, depending on the light levels.
Light Grasp
The light grasp of the eye as a function of aperture is quite modest. At a dark
adapted pupil diameter of about 6 mm, the pupil area allows only 3.5% the light
grasp of a 150 mm (6") telescope. However the eye is capable, under optimal
conditions, of perceiving a light source emitting only a few photons per second.
Thus, what it lacks in aperture it makes up in retinal sensitivity.
Resolution
The minimum angular interval that can be resolved by a normal human eye varies
with the type of resolution required, but is usually given as Rv = 70 to 140
arcseconds (0.00034 to 0.00068 radians). This is approximately 6 to 3 pairs of
alternating black/white lines within 1 millimeter viewed from 25cm (152 to 76 line
pairs per inch, viewed from 10 inches).
The next diagram describes the eye's optical resolution in the abstract, as the spot
diameter of a point light source on the retina as a function of pupil diameter,
based on calculations with the Navarro ideal or mathematical "eye".

The Navarro eye allows


the spot diameter to be
separated into the
contributions from the
size of the diffraction
artifact (the Airy disc,
which gets smaller with
larger pupil aperture)
and from optical
aberrations in the cornea
and lens (geometric
spreading, which
increases with aperture
larger than ~3 mm). At
the pupil sizes
characteristic of most
visual astronomy, the
spot diameter in a
normal eye is generally
far in excess of 25 m (0.025 mm); but note that a pupil diameter of around
3.5 mm would apply to most lunar and some planetary observing. Values will
generally be worse for older and visually impaired observers.
A common method for estimating the eye's peak visual acuity is to calculate the
average spacing between two cones in the fovea as the diameter of a single cone,
and to assume that visual resolution is equal to this average spacing (the gap
between two point stimuli is as wide as one cone). The average cone density is
estimated to be around 190,000 per mm2 in the foveola (diameter on the retina of
about 0.35 mm), and about 100,000 per mm2 in the fovea (diameter of about
1.85 mm). This works out to an average cone diameter of about 0.0025 mm in the
foveola and 0.0034 mm in the fovea, given the characteristic hexagonal "tiling" of
cone aperture.
With the eye's average internal focal length of 24mm, these widths define an
angular separation of between 0.006 to 0.0081, or 0.36 to 0.49 arcminutes.
However, the minimum spot diameter of the eye under optimal conditions
(diagram, above) is about 7 m or 0.007 mm, which is roughly the diameter of
three cones. Thus the cone spacing method for predicting visual acuity overstates
the eye's detection threshold under optimal conditions by a factor of roughly 3
and this is consistent with the measured performance acuity of the normal eye
under bright light, which is between 1 and 2 arcminutes across the fovea.

However, the spot


diameter under low
luminance adaptation is
well above 0.025 mm,
which implies a visual
acuity in most
observational astronomy
that is as poor as 10
arcminutes. This
degradation of visual
acuity with increasing
pupil aperture is clearly
indicated in the
modulation transfer
functions of normal eyes
at various pupil apertures,
measured by
photographing the spot
size actually imaged on
the retina of subjects
(diagram, below).
Finally, the task dependence of visual acuity is reflected in the different
discrimination thresholds that characterize different monocular visual tasks:
Shape recognition - Black on white letter shapes can be resolved (e.g.,
rotational orientations of the shape "E" can be identified) when elements of the
letter are at least 5 arcminutes wide.
Grating resolution - Black on white line gratings (the lines and the spaces
between them are of equal width) can be discriminated from a uniform gray
background when the spacing between black lines is at least 2 arcminutes.
Point recognition - Two black points on a white background can be resolved
when the space between them is at least 1 arcminute; white points on a black
background can be discriminated at about 2 arcminutes (due to spreading of the
image).
Vernier (Nonius) acuity - The misalignment between two black lines placed
end to end or the vertical misalignment of two dots can be perceived when the
lateral displacement is at least 10 arcseconds. This form of acuity is the principle
behind the heliometer and the reason why it was the tool that first measured
stellar parallax (Bessel and 61 Cygni).
Contrast Sensitivity
Contrast or modulation range

The troland is a measure of retinal illuminance equal to the illumination from a


surface emitting 1 lux per square meter into a pupil area of one square millimeter;
it increases both with an increase in source brightness and an increase in pupil
aperture. For comparison, 1000 trolands is approximately the brightness of a
"white" paper viewed under bright indoor lighting, and 0.1 trolands is
approximately the brightness of white paper viewed under moonlight.
Most "deep sky" astronomical targets present retinal illuminances of less than 0.1
trolands, which implies a visual acuity limited at around 2 to 5 cycles per degree or
Rv = 360 to 900 arcseconds.
Field of View
Because the eye's normal point is the standard for visual magnification, calculation
shows that the eye's true field of view is equal to its apparent field of view.
As a simple test of your true field of view: stand and look straight ahead with both
eyes, then extend your arms on either side, fingers facing forward. Wiggle your
fingers while moving both your arms forward and backward until you find the point
where the movement of the fingers of both hands is just visible on both sides of
your visual field. The angle between your arms will be about 120.
However, the fact that your fingers are barely visible at 120 means that your
useable field is much smaller. How small? At the minimum, the fovea is only able
to resolve the central ~1 of your visual field. Perceptual research also shows that
viewers will say a single object fills their visual field when it subtends about a 20
visual angle a generalization first made by Leonardo da Vinci. For example, the
industry recommended wide screen TV viewing distances produce a visual
angle to the wide screen diagonal of between 20 to 40, with the average around
25 an image width that is said to produce "an immersive feel". By that
standard, the 40 apparent field of traditional eyepieces would be fully immersive.
Currently available commercial eyepieces offer apparent fields of view between
40 to 100. The choice of one end of that range over the other seems to tip at an
apparent field of around 60, and can depend on several considerations: esthetic
preference (the "majesty factor" as opposed to "looking through a drinking
straw"), visual task (sweeping for comets as opposed to resolving double stars),
the telescope mounting and balance sensitivity (super wide field eyepieces can be
very heavy, especially at long eyepiece focal lengths; super wide field eyepieces
allow extended "drift" viewing of objects with hand guided Dobsonian altaz

mountings), objective relative aperture (off axis eyepiece aberrations are more
noticeable at smaller focal ratios) and price (super wide field eyepieces are
generally much more expensive than traditional designs).
Off axis image quality, in telescopes with focal ratios down to about /6, can be
superior in some super wide angle eyepieces than in some traditional eyepieces,
and image quality at the center of the field (on axis) is, for almost all oculars
operating at all focal ratios, optically perfect. Only at objective focal ratios below
/5 do optical considerations become important.
On the other hand, super wide field eyepieces require the observer to "look
around" in the wide available field in order to see clearly, and this movement of
the eye and head can cause vignetting or aberrations to appear. (Bias
disclosure: I prefer eyepieces with apparent fields no greater than 70, mainly to
avoid these effects.)
In general, preference for wide or "narrow" apparent field eyepieces is not an issue
of optical quality or the "immersive" visual area useful to the steady gaze of the
human eye, but depends on the astronomer's personal preferences in visual
esthetics, ergonomics and visual tasks, the convenience of use with a specific type
of telescope ... and price.
Focus Diopters
Diopters are used to measure the distance that a lens must be moved forward or
backward along the optical axis to produce a change in the image focus equal to
one diopter of the eye. This focus diopter is found as:
e2/1000 mm.
Thus the movement required to adjust the focus of a 20mm eyepiece by one
diopter is (202)/1000, or 0.4mm.
Focus diopters may be positive or negative. At zero focus diopter, an eyepiece
projects collimated light rays of light parallel to the optical axis, which can then
focused by a relaxed normal eye as if the light came from a far distant source.
Therefore the optimal focus diopter for a normal eye is zero or very slightly
negative.

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic
optical principles.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of
historically important eyepiece designs.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces,
and results from my collection.
Amateur Astronomer's Handbook by J.B. Sidgwick - excellent basic chapters in
optics, light gathering, resolution, magnification and more.

Telescopic Limiting Magnitudes by Bradley Schaefer - an attempt to predict


telescopic limiting magnitudes using visual data, star color, age of observer, sky
brightness and other factors.
Human Eye from Handbook of Optical Systems, Vol. 4: Survey of Optical
Instruments by Herbert Gross, Fritz Blechinger & Bertram Achtner (eds.). (Berlin,
DR: Wiley-VCH, 2008).
Visual Acuity - Summary of the various methods used to test the human visus.
Average optical performance of the human eye as a function of age in a
normal population by A Guirao, C Gonzlez, M Redondo, E Geraghty, S Norrby
and P Artal. Investigative Ophthalmology & Visual Science Jan. 1999, pp. 203-213.
The N.A.A. Telescope Calculator - handy and accurate calculator page for most
optical parameters of a telescope/eyepiece combination.
Eyepiece Focal Length Measurement - Jim Easterbrook explains how to
measure eyepiece focal lengths in grating projections.
The Collimation - an excellent optical discussion by Thierry Legault
What Is a MTF Curve? - Basics of the MTF.

Last revised 6/20/12 2012 Bruce MacEvoy

Exit Pupil
The exit window is the image
of the eyepiece field stop as
viewed through the eyepiece
(image, right). The exit pupil
is the image of the telescope
aperture stop viewed
through the eyepiece
(image, far right).
Although the exit pupil as
image is visible hovering
above or just inside the eye
lens of the eyepiece when viewed from an oblique angle, it is generally located at
the cross section of the telescope's afocal output where the projected, centered
aperture image appears physically smallest (as shown here). (This is called the
Ramsden disc in the United Kingdom, in honor of the 18th century optician who
first described it.) The exit pupil, as Ramsden disc, is the point where the
telescope afocal image is most concentrated and therefore brightest.
A commonly suggested procedure to measure the exit pupil always stated in
millimeters is with a reticule magnifier. This sometimes precedes the use of the
exit pupil in the calculation of system magnification as the ratio between objective
aperture (Do) and exit pupil diameter (de) that is, as the inverse of the beam
compression of the system:
[1] Mt = Do/de
However, measuring the size of the exit pupil is inconvenient, and it is simpler and
more intuitive to calculate magnification as the ratio of objective focal length (o)
to eyepiece focal length (e). Then the exit pupil diameter can be calculated as the
ratio between the objective aperture and the magnification produced by the
objective/eyepiece combination:
[2] de = Do/Mt
This seems to define the exit pupil is the ratio of aperture to magnification: an exit
pupil of 1.0mm means the aperture (in millimeters) is equal to the magnification;
an exit pupil of 2.5mm means the aperture is 2.5 times the magnification, and so
on. But that is a misinterpretation. In fact, this definition of the exit pupil is the
same proportion that defines the telescope image illuminance in relation to the
retinal illuminance of the naked eye:
It = (Do/Mt)2, therefore de = It / 2
which means the exit pupil is effectively
image illuminance standardized on pupil
aperture. And if the eye pupil is assumed
to be a personal constant then the exit
pupil is proportional to the perceived
image brightness.
The exit pupil is optically the smallest
aperture through which all the afocal light
from the telescope must pass, its diameter

is related to the luminance of the telescope image in the same way that the
opening in a camera diaphragm determines the amount of light that can enter a
camera (graph, left). However, the rising curve on the left side is accompanied by
an increasing true field of view and decreasing image magnification, which also
brighten the image. The "shutter" and magnification effects combine to brighten or
darken the image in inverse square proportion, and the difference in the
illuminance of the telescope image equals the ratio of two exit pupils squared:
[3] L = (de.2/de.1)2
The peak image brightness occurs when de = . At that point, the telescopic image
is exactly as bright as the naked eye image of the same area of sky (omitting
transmission losses in the instrument).
Because any larger exit pupil diameter will exceed the nominal eye pupil aperture:
de > . In this situation, the fixed eye pupil is the limiting factor. As the exit pupil
becomes larger, the image luminance declines because the area sampled by the
fixed pupil size is a smaller part of the total exit pupil area, effectively stopping
down the objective aperture.
However, if we substitute into formula [2] the alternative definition of
magnification (as o/e), we find a third definition of the exit pupil as the ratio
between the eyepiece focal length and the telescope relative aperture (No):
[4] de = Do/(o/e) = e/(o/Do) = e/No
An exit pupil of 1.0 mm is therefore produced by an eyepiece focal length that
equals the telescope relative aperture, a 3 mm exit pupil is produced by an
eyepiece focal length 3 times the telescope relative aperture, and so on. The exit
pupil is simply the eyepiece focal length projected by multiplication with a constant
value the relative aperture.
The value at which all elements form equalities is an exit pupil of 1.0:
de = 1.0, Mt = Do, e = No
where Do and e are in millimeters. Changing the exit pupil (which normally means
either changing the eyepiece e or stopping down the aperture Do) involves
reciprocal changes in the other two equalities, which leads to the general form of
exit pupil relationships:
[5] de = n, Mt = Do/n, e = nNo
The exit pupil by itself gives no information about the optical system, other than
the diameter of the Ramsden disc and the image brightness relative to the naked
eye brightness [as (de/)2]. The diagram (below) illustrates these exit pupil
calculations.

Three fixed physical quantities of the telescope system (objective focal length o,
eyepiece focal length e and objective aperture Do, shown in white, above) are
used to calcuate two system ratios (relative aperture No and magnification Mt,
both in orange). The fourth physical quantity, eye pupil , is assumed constant
and is only necessary to benchmark relative image brightness.
So long as the eyepiece focal length and system focal ratio remain constant, all
other physical parameters related to the telescope objective the aperture, focal
length, magnification, resolution limit and image scale can be changed without
changing the exit pupil.
Note specifically that the exit pupil has no fixed relationship to magnification: an
exit pupil of 2.0 mm can equal a magnification of 125x with a 250 mm (10")
aperture or a magnification of 40x with an 80 mm (3.2") aperture. As with the
eyepiece itself, the magnification produced by an exit pupil depends on the focal
length of the objective it is used with.
Since eyepieces are made to be swapped in an out of different optical systems, the
exit pupil is commonly used to choose an eyepiece magnification that is suitable
for the parameters of any specific telescope system it's easy math to multiply
focal ratio by the desired exit pupil value to get the necessary eyepiece focal
length. Handy in itself when using many different telescope systems, this
procedure is sometimes misapplied to the factitious twin "problems" of wasted
light (or wasted aperture) and empty magnification.
The astronomer prevents "wasted light" by using magnification that creates an exit
pupil that is not larger than the eye pupil. As explained above, the "problem" is
that an eye pupil smaller than the exit pupil will block some light from entering the
eye, effectively "stopping down" the objective aperture. But with most
commercially available astronomical equipment, it is difficult to exceed the eye
pupil aperture. With an /6 telescope and a pupil aperture of 6 mm (my measured
dark adapted pupil size), an eyepiece with focal length greater than 36 mm would
be necessary. Even in an /4 telescope, only eyepieces above 24mm would cause
the problem. And there is no "problem" if a longer focal length eyepiece is chosen
to give a wider, lower power view of an extended object. Then low magnification is
used to increase the true field of view at the "cost" of reduced image luminance, in
the same way that high magnification is used to increase visual resolution at the
"cost" of a smaller true field of view.
At the other extreme, "empty magnification" is said to be any image enlargement
greater than necessary to make the Airy disc just visible to the eye. Since the
linear width of the Airy disc is dependent only on the relative aperture (as
2.44No), it is always 0.00135No radians wide. This can be rewritten as the Airy
disc angular diameter [(0.00135 x 206265)No = 279No] as a ratio of the exit pupil:
s = 279/de arcseconds
Alternately, given your personal visual resolution limit of Rv (values cited in the
optical literature range from 70" to 280" or more), the corresponding exit pupil at
which the Airy disc will become visible is
[6] de = 279/Rv mm
which motivates the conclusion that an exit pupil de < 0.5 mm only provides
"empty magnification" because the Airy disc is then more than twice the size of the
most conservative resolution limit (280"). In fact, adopting the typical visual

resolution threshold (120") indicates diffraction artifacts are visible when the exit
pupil is as large as 2.3. But these abstract considerations ignore the complicating
factors of image motion (due to atmospheric turbulence), brightness and contrast,
all strongly affected by magnification. Exit pupils below 0.2 are routinely valuable
when resolving close, bright double stars, and William Herschel employed exit
pupils as small as 0.03 when studying double stars with his "most excellent" /13
reflector. There is no such thing as "empty magnification" only magnification
that is not optimal for the visual task at hand.
However the physical diameter of the Ramsden disc does have important visual
consequences. A small exit pupil (less than 1.0 mm) projects somewhat like a
pinhole, which can cause vivid shadowing of suspended internal debris or "floaters"
on the retina. At the same time the tiny Ramsden disc must be centered in the
pupil to produce a view of the entire eyepiece field, which becomes a kinesthetic
challenge in very wide angle eyepieces or awkward viewing postures unless the
observer has a chair or ladder for support. On the other hand, a smaller exit pupil
is only refracted by the central portion of the eye lens and cornea, minimizing the
effects of astigmatism. For these reasons alone, many older viewers prefer exit
pupils between 1-2 mm.
More significant problems with the exit pupil occur when observing during daylight
with a binocular or spotting telescope. Daylight images will contract the pupil
aperture to 2 mm or less, and this makes alignment of the eye pupil and exit
pupil(s) more difficult. It also makes any spherical aberration of the exit pupil
more noticeable, and minimizes aberrations produced by the observer's cornea
and lens.
Resolution
The resolution of an optical system is measured as the smallest angular width that
the system can image at the focal surface. In astronomical systems this limit is
determined by three things: the quality of the optics of the system, the turbulence
or distortion introduced by the atmosphere, and the quantum wave properties of
light. Only the last property can be calculated directly as the reciprocal of aperture,
1/Do.
The Diffraction Artifact. Analysis of the aperture resolution limit depends on a
quantum (physical) rather than geometrical analysis of light. If the light
wavefronts from a distant star are traveling in direction x parallel to the optical
axis of a telescope, then the width of each wavefront is measured along a
dimension y perpendicular to the direction of light. Because the star is so far away,
the scale of the y dimension is effectively infinite.
If we specify a specific energy of light by wavelength, such as the commonly used
= 555 nm (the average wavelength of a normal eye's peak photopic sensitivity),
then the quantum indeterminancy of the location and momentum of any photon in
the wavefront is distributed across the entire dimension y, and both the energy
and location of the photon can have a precise value, which means it can carry
information about the emitting source of the photons.
However, once the wavefront passes through the aperture stop of a telescope, all
the indeterminacy in the y dimension must be reduced to the aperture width D.
We require a fixed momentum and wavefront location for each photon at the focal
surface of the objective, because the photon contributes to the formation of an
image. But this fixed position must also include the indeterminacy contained in the
y dimension of the wavefront, which produces an "error" or uncertainty in the

specific location of the photon in the image. This quantum uncertainty is defined as
no smaller than:
U = h
where h is Planck's constant, 6.626 x 10-34, and is the frequency (energy) of the
light, and where the speed of light c is their product, c = . Casting these
quantum relationships in a form that defines the indeterminacy of a circular
aperture of diameter D yields:
UD = hc / hD = h / hD = / D
Thus, for a "green" wavelength of 555 nm in a 10" (254 mm) telescope, the
angular width of the indeterminacy becomes:
UD = 0.000555 / 254 = 2.185 x 10-6 radians
In the light from a far distant star, the result of the cumulative uncertainty from all
photons focused to a single image point does not have a simple "fuzzy" image
structure around the point, due to the destruction or reinforcement of light waves
as different parts of the wavefront are superimposed at the same location on the
image plane. Instead it appears as a central concentration of light surrounded by
one or more rings, the center and rings separated by concentric dark intervals.
The central concentration is called the Airy disc (after the Cambridge professor
George Biddell Airy who first described it mathematically in 1835), surrounded by
concentric diffraction rings. The angular dimension of these features, in relative
units of /D, the relative luminance (as a proportion of the peak Airy disc
luminance) and the percentage of total light energy enclosed by each dark ring are
given in the table (below).

dimension (k) in /D radians


feature

peak
percent of
illuminance total light

radius of
maximum

diameter of
maximum

Airy disc

1.0

first gap

1.22

2.44

83.8%

first ring

1.64

3.28

0.017

second gap

2.24

4.48

91.0%

second ring

2.66

5.32

0.0041

third gap

3.24

6.48

93.8%

third ring

3.90

7.80

0.0016

fourth gap

4.24

8.48

95.3%

the Airy disc and diffraction rings


of a "point" light source

Objective Resolution. The diffraction


artifact of Airy disc and encircling rings
is the smallest possible image of a
"point" light source. As a result, the
diffraction artifacts created by two
"incoherent" or physically unrelated
light sources (such as two stars)
separated by an arbitrarily small
angular distance will appear to merge,
because the angular separation
between them is smaller than the
angular diameter of the artifacts
produced by that aperture.
However, the spacing between two
diffraction artifacts, standardized on
the /Do radian width characteristic of
every aperture, can be scaled to
represent the minimum separation
between two point sources necessary
to clearly resolve or separate their
diffraction artifacts. This is the basis
for an angular resolution limit for point
light sources given the aperture Do.
In the Rayleigh resolution limit (diagram, right), the required spacing places the
peak intensity of each Airy disc over the first dark gap of the other star, which is
given as k = 1.22 or 1.22/D radians in the table above. This produces a merged
minimum brightness betwee the two stars that is roughly 70% of the peak
brightness of each Airy disc. The three most commonly used resolution limits in
astronomy are:
RR = 1.22/Do206265 arcseconds (Rayleigh criterion)
RD = 1.02/Do206265 arcseconds (Dawes criterion also approximately the Airy
disc full width half maximum or FWHM)
RS = 0.95/Do206265 arcseconds (Sparrow criterion)
where the factor 206265 is used to convert radians into arcseconds, Do is in
millimeters, and the wavelength of light is usually chosen to match the peak foveal
response, which is distributed over a rather broad range between approximately
530 to 580 nm.
Note that the more stringent Sparrow and Dawes resolution criteria also have a
more limited application. Both apply only to two stars of equal brightness viewed
through a circular telescope aperture with no central obstruction. In addition, the
Dawes resolution criterion is specified for matching 6th magnitude stars, and the
Sparrow criterion becomes larger as the magnitude difference between the two
stars increases. In contrast, the Rayleigh criterion is not affected by the absolute
or relative magnitude of the two stars and is not affected by any central
obstruction, and for those reasons is the criterion most generally used.
Adopting arcsecond measurement units and a reference wavelength of =
0.000555 mm, the more convenient formulas to calculate the resolution limits (in
aperture millimeters) are:

Ro = 140/Do (Rayleigh)
Ro = 117/Do (Dawes)
Ro = 109/Do (Sparrow)
and in aperture inches:
Ro = 5.50/Do (Rayleigh)
Ro = 4.60/Do (Dawes)
Ro = 4.28/Do (Sparrow).
The factors used in these formulas depend on the wavelength of light used as
reference. Thus, the Dawes criterion is more commonly given as 116/Do or
4.56/Do, which has been calculated on a wavelength of = 550 nm. These
differences are in practical application inconsequential.
In an afocal system with circular aperture, the unweighted quantum formula is
used to compute the minimum resolvable interval between two lines:
o = /Do206265 arcseconds
Thus, a diffraction limited 10" (254 mm) telescope can theoretically resolve two
lines separated by about 0.45 arcseconds, but two point sources (according to the
Rayleigh criterion) separated by about 0.55 arcseconds. "Theoretically" means
optical quality matters, and atmospheric turbulence intervenes.
Visual Resolution (With Eyepiece). The resolution delivered by the objective to the
image surface is only one part of the total optical system. Visual recognition of the
artifact, or any dimensional feature on the image surface, depends on its image
size relative to the resolution of the eyepiece/eye combination.
Unlike the angular width, the linear dimension of the diffraction artifact on the
image plane scales only with relative aperture, independent of aperture and focal
length:
o = kNo millimeters
where k is the dimensional of the feature radius or diameter taken from the table
above. For example, the linear diameter of the Airy disc (at = 0.000555 mm) in
an /8 telescope is 2.44 x 0.000555 x 8 = 0.011 mm.
This diffraction limited image dimension must be compared to the minimum
physical width that can be separated by the observer's eye with the assistance of a
specific eyepiece, defined as:
Re = eRv/206265 millimeters
where Rv is the visual resolution limit (in arcseconds, converted to radians by
division by 206265) and e is the eyepiece focal length (in millimeters). Thus the
minimum width that can be resolved with a 10 mm eyepiece, given a visual
resolution limit of Rv = 120 arcseconds, is 10 mm x 120"/206265 = 0.0058 mm.
Comparison of the two widths shows that the linear diameter of the Airy disk
(0.011 mm) is roughly twice the width of the resolution limit (0.0058 mm), and
therefore would be easily visible in an /8 optical system with a 10 mm eyepiece.
However this might not be the case if the eye were fully dark adapted and the

pupil aperture larger.


Appearance of the Diffraction Artifact.
Given a fixed aperture and focal ratio,
optical theory stipulates that
regardless of the magnitude of the
"point source" the diffraction artifact
will always have the same linear and
angular diameter and the same
proportional illuminances between the
Airy disc and any of the rings around it.
Remarkably, this does not correspond to the detailed visual appearance.
First, as Sidgwick (p.39) explains, "At normally encountered /ratios (say, /5 to
/20) no intensity gradient across the disc is perceptible, the border between the
disc and the first minimum appears nearly sharp, and the rings are brighter than
theory would indicate, the first ring being not much fainter than the disc itself. ...
The visible extent of the disc, like the number of rings visible, varies for a given
instrument with the brightness of the source, although the discs are in fact the
same size, irrespective of brightness."
The paradox of the varying disc diameter and number of rings was explained by
Airy as due to the luminance threshold of the eye. This is illustrated by plotting the
diffraction artifact using a log scale for the luminance (diagram, right), since the
relative brightness of lights viewed against a dark background scales as
b = b2/b1 = 100.4(m2-m1)
The dotted horizontal lines in the diagram show the relative effect of a visual
threshold that extends down to either ~1/10th to ~1/1000th the peak Airy disc
brightness. If a magnitude 2 star is bright enough to present three rings, this
explanation requires the third ring of a mag. 2 star to appear as bright as a
magnitude 9.5 star
(diagram below, A).
A threshold explanation for
the changing angular size
of the disc does not explain
the fact that the first
diffraction ring contracts
around this decreasing
disc, becoming slightly
thinner and separated from the disc by a constant or slightly narrower first gap
(diagram right, C). It does not keep the same diameter regardless of the star
magnitude, which would produce in very faint stars a visibly enlarged first gap
around a much reduced Airy disc (diagram right, B).
Magnification
Magnification is the increase in the angular size (apparent width) of an object in
comparison to the naked eye size of the object. In telescope optics, magnification
results from the combined effect of the telescope objective and eyepiece, with the
optics of the eye taken as the reference.
The eye has a variable focal length: it can focus objects at different distances
within the fixed distance from the cornea and lens to the retina. Which focal length

provides the reference? We use the distance that provides the maximum image
size on the retina, the minimum object distance at which an object will appear
comfortably in focus at full accommodation. For the average human observer this
near point is 250 mm (25 cm).

An object moved closer to the eye than the near point will appear larger, but also
out of focus. Focusing the eye on an object closer than 250 mm is made possible
by a magnifier lens or eyepiece.
The magnification is then the ratio between the angular size of the magnified
object and its angular size as viewed by the naked eye at the near point distance
of 250 mm. The distance size rule states that the ratio between the two distances
is the inverse of the ratio between the two visual angles. Thus, an object imaged
at 50 mm will appear magnified by 250/50 = 5 times, and will subtend a visual
angle tan() = 5 x tan().
The eyepiece effective focal length (EFL) is determined as the distance between
the eyepiece focal plane and the eyepiece principal plane, which is located where
the field angle of the exit window intercepts the radius of the field stop. This
distance must be divided into the visual angle of the field stop as it appears at
250 mm, so the eyepiece magnification (diagram below, right) is calculated as:
Me = 250/e
If the ocular is used as an actual magnifier, with the object held at the focal plane
of the eyepiece, then its magnification is 250/e+1.

The angle of the principal ray from an object through the telescope objective
remains constant, but this angle projects a greater field height at a longer focal
length. Therefore the telescope objective magnification is determined by its focal
length. Now, what telescope focal length will produce an image of a distant object
(such as the moon) that, when the image is viewed at the near point, will appear
exactly the same size as the naked eye image? The answer is: when o =
250 mm. Therefore the telescope magnification (diagram above, left) is equal to:
Mo = o/250

Because one magnification has a multiplying effect on the other, the magnification
of the combined telescope and eyepiece system is the product of the separate
telescope and eyepiece magnifications. The factor 250 cancels out, leaving the
ratio of the two focal lengths:
Mt = MoMe = o/250 x 250/e = o/e
Thus, in a 10" (254 mm) /8 telescope with a 20mm eyepiece, the magnification
is:
Mt = (254 x 8)/250 x 250/20 = 8.13 x 12.5 = 101.6
When caclulating magnification, keep in mind that manufacturer supplied data on
telescope and eyepiece focal lengths is often inexact, and therefore magnifications
calculated from them should be considered approximate. Magnifications greater
than about 10x (e < 25 mm) and free of serious aberrations are not possible with
a single lens; a compound lens is required.
Magnification affects the image illuminance: as magnification increases, a smaller
part of the sky (the true field of view) is passed to the eye, and therefore a
smaller quantity of light is being used to fill the apparent field of the eyepiece. In
other words, as magnification increases, the image becomes dimmer. These
effects can be calculated starting with the exit pupil equality de = Do/Mt = ,
assuming an eye pupil nominal diameter of 6 mm (table, below).

image
brightness

aperture (mm)
de
100 150 200

(naked eye = 1)

250

300

350

400

500

1.0

17

25

33

42

50

58

67

83

0.1

1.9

54

79

105

132

158

183

211

263

0.01

0.6 167 250

333

417

500

583

667

833

0.001

0.19 528 790 1053 1317 1580 1830 2107 2633

The telescopic image of an extended object (the Moon) can never be brighter than
the naked eye image of the same celestial area, due to "stopping down" of larger
beams by the eye pupil. The telescopic image of a star brightens with aperture,
because the larger quantity of light gathered by the larger aperture is concentrated
into a star image that does not increase in size (it remains a "point" source of
light).
Magnification Selection
Various guidelines have been proposed to determine the optimal or most useful
magnification to use with a telescope. These are generally stated as multiples of
the aperture (in millimeters or inches), as these determine the brightness (light
grasp) and detail (angular resolution) of the image to be examined: and
brighter, more detailed images respond better to high magnification.
The benefits of higher magnification are darkened sky background and better
visibility of faint stars, easier visual detection of image features at the resolution
limit, smaller field curvature and reduced optical aberrations. The drawbacks are a

smaller true field of view, decreased image luminance, amplified atmospheric


turbulence, exaggerated instrument vibrations, and greater difficulty centering and
manually tracking a target object. The pointlike projection of a tiny exit pupil can
also silhouette "floaters" in the eye and dust on the eyepiece.
The magnification necessary to visualize a given spatial frequency is the angular
resolution of the eye (Rv) divided by the resolution of the telescope system
(calculated above):
Mmin = Rv/Ro
where Rv is the minimum angular width that can be resolved by the naked eye,
quoted (depending on context) as ranging from 60" to 120" up to 360", depending
on the eye pupil diameter and the type of resolution required. This optimal
magnification range is usually supplemented with lower and higher powers, and
can be stated as simple multiples of aperture (for magnification) and reciprocal
multiples of relative aperture (for the corresponding eyepiece focal length). The
five common recommendations and the justifications offered for them are:
Mt = Do/, e = N Normal magnification This defines the magnification at
which the exit pupil equals the pupil aperture of the observer. The rationale is that
any lower magnification produces an exit pupil that is larger than the eye pupil,
which effectively stops down (reduces the light grasp) of the objective. But this is
a dubious rule. Few observers have actually measured their dark adapted pupil
aperture, and the customary value (8mm) is typically too large. (Mine is 5.9mm.)
The main objection is that low magnification has benefits unrelated to light grasp:
the true field of view is larger, pointing and tracking become easier and less
critical, and the effects of poor seeing are minimized. Finally, observing bright
objects such as the Moon can contract the pupil to 3mm or less and, with
commercially available eyepiece focal lengths (e 50mm), telescopes at relative
apertures above /6.3 cannot reach the prescribed "normal magnification"
anyway. A more practical guideline is:
Mt = 0.25Do to 0.33Do, e = 3N to 4N Minimum useful magnification The
lower bound at this magnification level is not determined by the diameter of the
eye pupil, but by the angular diameter of the target (star cluster, extended nebula,
etc.) and by the relative aperture of the telescope, which can introduce distracting
aberrations in wide field eyepieces at low powers. Note that magnification between
0.25Do and 0.5Do is useful for deep sky observing when a turbulent atmosphere
precludes detailed observation of planets or the Moon.
Mt = 0.5Do, e = 2.0N Minimum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the optimistic minimum angular resolution of the eye (i.e.,
~1 arcminute, or 60/Ro). Obviously, the quality of the observer's eyes affect this
rule significantly.
Mt = 1.0Do, e = 1.0N Maximum optimal magnification When the
magnification causes the smallest (diffraction limited) image detail to be just
visible given the conservative minimum angular resolution of the eye (i.e., ~2
arcminutes, or 120/Ro). This is equivalent to "Whittaker's rule", usually stated as
Mt = 25Do when aperture is in inches (25 is just the number of millimeters in an
inch).
The "optimal magnification" rules are based on visual resolution limits, but it is

useful to observe with eyepieces in both focal lengths because an increase in


magnification causes sky brightness to decrease and very faint stars to become
more visible, despite the increase this may cause in the visible effects of
atmospheric turbulence.
Mt = 2.0Do to 4.0Do, e = 0.5N to 0.25N Maximum useful magnification
When the magnification causes the smallest (diffraction limited) image detail to be
just visible given the minimum angular resolution of the eye with a fully dark
adapted pupil (~5 arcminutes, or 300/Ro). This rule is highly dependent on
observing conditions and the observing target. Planetary and lunar observers will
generally prefer a lower magnification than double star observers, because
planetary details are low contrast and easily masked by poor seeing, whereas
stellar Airy discs tolerate high magnification very well, even under poor seeing.
In all cases, the maximum useful magnification has been exceeded when the
image is too dark (due to the magnification of the image), too blurry (due to the
quality of the optics, to atmospheric turbulence), or too difficult to view (due to
"floaters" in the observer's eye, small eye relief, etc.).
For the visual astronomer there is a less arbitrary approach to eyepiece selection.
Using a symmetrically cut aperture stop, shown here, a telescope trained on a
bright star will present a series of parallel bright and dark bands, spaced according
to the Rayleigh criterion (140/Do). View the bands with various eyepieces until you
find the eyepiece that yields a minimum but confident detection threshold. Divide
the telescope focal length by the eyepiece focal length to get the magnification,
then multiply the magnification by your telescope's Rayleigh resolution limit. This
yields your visual resolution threshold in arcseconds. (Mine is about 100.)
The simplest and most reliable rule is: use the magnification that produces the
most useful image. Atmospheric turbulence, the magnitude of stars or the surface
brightness and surface contrast of extended objects, the sky brightness, the type
and optical quality of the instrument, the quality of the observer's eyes and the
specific visual task or pastime all contribute to the choice of the optimal and most
useful magnification.
The corollary rule is: change the eyepiece often. No matter what the visual task or
observing target, trying different eyepieces will verify the best choice, and can
reveal unexpected details in the peripheral field or the target. Magnification
guidelines should be used merely guide the range of eyepieces selected for the
occasion.
Apparent Field of View
The apparent field of view (AFOV) of an optical system is the image of the field
stop in the image space, called the exit window. The AFOV is measured as the field
angle of the radius () or diameter (2) of the field edge or image limit (diagram,
below). The tangent of this angle is equal to the radius of the field stop opening
divided by the eyepiece back focal length (BFL).
The physical diameter of the eyepiece barrel and location of the field stop are
dictated by the optics of the eyepiece (its abaxial aberrations and design AFOV)
and the ratio of the objective (see here). The field stop excludes abaxial light
from the objective that would produce aberrations in the eyepiece image or an
unacceptably dark image, and blocks stray light reflected from the interior of the
telescope tube. The barrel diameter limits the maximum internal diameter of the
field stop, to ~27 mm in a 1.25" eyepiece barrel and ~48 mm in a standard 2"

eyepiece barrel. In all well made eyepieces, the field stop will define the apparent
field of view and will be crisply in focus.

The apparent field of view of an eyepiece is normally specified by the


manufacturer, along with the eyepiece focal length and design type. However,
the nominal AFOV of an eyepiece is often inflated due to the effects of positive
distortion, especially in "wide field" eyepieces (AFOV > 60).
When the manufacturer's apparent field specification appears inaccurate, or is not
known, then the AFOV can be measured in three ways:
(1) Projective Measurement. Support the eyepiece so that it is lying on its side
with the eye lens about 2 feet from a smooth wall or flat surface. (The seam
between pages of an open hardcover book, or between two books of the same
thickness laid spine to spine, is convenient.) Support a high powered flashlight in a
similar way so that it shines into the field lens (barrel end of the eyepiece) from a
distance of a few feet. (The eyepiece and flashlight should be lying perfectly level,
the optical axis of the eyepiece and flashlight must be perpendicular to the wall,
and the flashlight must be "collimated" or centered on and parallel to the eyepiece
optical axis, which can be arranged by centering the shadow of the flashlight bulb
inside the projected circle of light.) The eyepiece will project the light from the
flashlight onto the wall as a circle of light larger than the flashlight beam. Project
the image onto a small white card held close to the eye lens, and move the card
back and forth until the projected circle appears smallest: this is the location of the
exit pupil. Measure the greatest width of the circle (D) and the projection distance
(P) from the eyepiece exit pupil to the wall. Then:
AFOV = 2arctan(D/2P).
For example, if the imaged diameter is 82.3 cm, and the wall to exit pupil distance
is 73.3 cm, then:
AFOV = 2arctan(D/2P) = 2 x arctan(82.3/(2 x 73.3)) = 2 x 29.3 = 58.6.
(2) Visual Measurement. Tape to a convenient wall a long strip of paper or tape
measure on which small increments (inches or centimeters) are clearly marked
and visible from a distance (Dm) of about 100 cm. Hold the eyepiece over one eye
so that the full field is clearly visible and superimposed over the measurement
tape, visible with the other eye. Manipulate the eyepiece position until the tape
appears to measure the full diameter of the eyepiece field (De). Then:
AFOV = 2arctan(0.5De/Dm).

For example, viewing a metric measuring tape from 100 cm, you measure an
apparent width of the eyepiece field as 93 cm. Then:
AFOV = 2arctan(0.5 x 93/100) = 2arctan(0.47) = 2 x 25 = 50
(3) Physical Measurement. Use calipers to measure the interior diameter of the
field stop, assuming it exists and can be recognized. If in doubt, look through the
eyepiece and use the tip of a toothpick or pencil to locate the field stop: this is
either the end of the eyepiece barrel or the interior edge of the lock nut holding
the field lens. (Note that many "super wide field" eyepieces locate the field stop
inside the eyepiece where it cannot be measured.) Then the apparent field is
simply the visual angle that the field stop radius would subtend at 250 mm,
divided by the apparent field radius . The distance size equation then makes this
equal to:
AFOV = 2arctan(0.5Dfs/e).
For example, if the field stop diameter is 23mm and the eyepiece focal length is
20mm, then:
AFOV = 2arctan(0.5Dfs/e) = 2arctan(11.5/20) = 2arctan(0.58) = 2 x 29.9 =
59.8.
True Field of View
The exit window admits to the eye an image that has been magnified by both the
telescope and eyepiece. The actual size of the exit window, if the image area it
defines were viewed with the naked eye, would appear much smaller and much
farther away (diagram, above). This is the true field of view (TFOV) of the
objective/eyepiece combination.
Standard Formulas. The TFOV is equal to twice the maximum field radius admitted
to the image plane by the field stop:
TFOVo = 2max = 2h'max/o.
The TFOV is equal to the apparent field of view (AFOV) divided by the
magnification:
TFOV = AFOV/Mt.
Thus, if the apparent field of view of an eyepiece is 50, and the magnification of
the telescope and eyepiece combination is 200x, then the true field of view is:
TFOV = 50/200 = 0.25; or 0.25 x 60 = 15 arcminutes.
In the diagram (above), an eyepiece with a 50 field of view offers the same visual
area as a circular opening 23 cm in diameter held perpendicular to the eye at 25
cm distance. Then true field of view is equivalent to the area of sky seen through
the opening at a distance of 50 meters (164 feet); multiplication by 60 converts
degrees to arcminutes.
If the field stop diameter and objective focal length are known, then the TFOV can
be calculated as:

TFOV = Dfs/o x 57.3.


This formula also shows that the maximum true field of view possible with a 1.25"
eyepiece barrel or a visual back of a certain diameter. For a 10" /8 telescope, the
maximum true field of view is:
TFOV = 27/2032 x 57.3 = 0.76 x 60 = 46 arcminutes.
Star Drift Measurement. If the field stop diameter and apparent field of view are
not known, the true field of view can be measured by means of star transit times.
This requires at least three measurements of the time (in seconds) it takes a
bright star, placed outside the eyepiece field of view, to appear at one side of the
eyepiece field, cross the center of the field, and disappear at the opposite side.
Care must be taken to hold the head in a fixed position, so that the entire
circumference of the field edge is visible, and to pass the star through the center
of the eyepiece field.
Then the true field is the average of the three timings multiplied by the cosine of
the star's declination (stars closer to the celestial pole will require a longer time to
transit the eyepiece field):
TFOV = 0.25cos(Declstar) x t.
The factor 0.25 is necessary to convert from seconds in time to arcminutes of
angular width. For example, using Regulus (alpha Leonis) as the transit star
(Declination = +1158'), three eyepiece transit times are 134, 131 and 138
seconds. Then:
TFOV = 0.25cos(11.97)(134+131+138)/3 = 0.25 x 0.978 x 134.3 = 32.8
arcminutes.
Given the TFOV (in arcminutes) and system magnification, the AFOV (in degrees)
is:
AFOV = TFOV / 60Mt.
The TFOV determined by star transit measurements will typically not correspond
exactly to the TFOV calculated from the manufacturer supplied eyepiece apparent
field of view, and the magnification calculated from the manufacturer supplied
eyepiece focal length. This is usually because the eyepiece AFOV and e are
nominal, inferred from the computer optical design rather than the physical
construction of the eyepiece; and because timing star transits includes some
measurement error.
Effect of Central Obstruction
Resolution is affected by wave diffraction effects, but so too is the potential of the
objective to transmit contrast information at different spatial frequencies.
Transmission is specifically affected by a central obstruction, such as the
secondary mirror in a Newtonian or Cassegrainian design, which creates an
annular rather than circular aperture geometry and diffraction at four edges on the
diameter rather than just two. This obstruction ratio is defined as:
obstruction ratio () = ds / Do

where ds is the diameter of the secondary obstruction, the mounting for the
secondary mirror.
The standard method to evaluate optical performance across a range of spatial
frequencies is the modulation transfer function or MTF. The MTF shows the
proportion of luminance contrast that is retained in the image formed by an optical
system as the spatial frequency of the luminance variations is increased.
Geometrically, the MTF calculation is equivalent to the area of overlap between two
aperture superimposed at a displacement equal to the spatial frequency.

The stimulus used is either a pattern of crisp black and white lines (to evaluate the
transmission of sharp edges) or a sine wave pattern of modulated grays (to
evaluate the transmission of tonal gradients). Both methods converge on similar
conclusions.
Modulation transfer is defined as the reduction in image contrast between the
range of the luminance in the image (LI) as a proportion of the range of luminance
in the target or stimulus (LS):
contrast ratio = (LImax LImin) / (LSmax LSmin)
Because diffraction produces a tiny amount of smearing in the image, contrast is
imperceptibly reduced at low spatial frequencies (wide spacing) but becomes
severe at high spatial frequencies (very narrow spacing). So the MTF is scaled to
zoom in on the high spatial frequences, defined as the inverse of the angular
spacing in multiples of the minimum theoretically resolvable angular spacing
(/Dmm). For example, the Rayleigh criterion of resolution is defined as 1.22/D,
so its spatial frequency in an MTF is 1/1.22, or 0.82. (The MTF is sometimes
graphed in terms of line pairs per millimeter (lp/mm) for a specific telescope
relative aperture, but this produces the same focus on high frequency spacing and
the compression of low frequency spacing at the far left of the graph.)

The green line in the


generic MTF diagram
(right) shows the curve for
an optically perfect and
unobstructed circular
aperture transmitting a
sine wave pattern with a
spacing frequency up to
the resolution limit
which in a 250mm /10
telescope is a stimulus
angular separation of 0.45
arcseconds or a spacing on
the image plane of
0.0055 mm or 182 line
pairs per millimeter.
The yellow line shows the
effect on the contrast ratio
of a 25% obstruction,
typically the upper limit in
a Newtonian reflector, and
a 50% obstruction, the
upper limit in a Cassegrain derived design optimized for astrophotography (Ritchey
Chrtien or Schmidt Cassegrain). Both patterns can be summarized (reading the
graph left to right) as (1) a negligible (less than 10%) contrast reduction in spatial
frequencies from the full image diameter down to about 10 times the Rayleigh
limit, (2) a significant drop in contrast resolution down to about 2 or 3 times the
Rayleigh limit, (3) an increase in resolution that can exceed an unobstructed
aperture by a small amount (due to enhanced contrast caused by overlapping
wave reinforcement), and (4) resolution matching the theoretical minimum at the
Rayleigh limit.
Using the 50% obstruction curve as an example, the minimum resolution occurs at
around 3 times the minimum resolution, or an angular width of about 1.5
arcseconds. This means the most severe image degradation will occur in lunar
features that are about 3 kilometers wide, but will produce a slight crispening in
features half that size. It will most affect features that are about 3% of the width
of Jupiter's disk at opposition or about 6% the width of Mars's disk at opposition.
Contrast on the Galilean satellites of Jupiter and on the disks of Mercury, Uranus
and Neptune would be strongly obscured or obliterated. Double stars separated by
about 1.3 arcseconds would be most compromised, but very close pairs would be
easier to separate.
Obviously, whether and how much a central obstruction matters depends
somewhat on the size of the aperture, but more on the specific task the telescope
is used to perform. Astrophographers can be tolerant of a large central obstruction
because they typically work on an image scale of 1 pixel = 5x the resolution limit,
and image contrast can be enhanced with image stacking and digital manipulation.
For visual astronomers the problem is more complex. Inherently high contrast
targets such as the moon and double stars will display relatively minor contrast
reduction, and the effect will depend on magnification, because human contrast
sensitivity peaks at around 7 cycles per degree (around 10 arcminutes) and
declines as frequency increases. At the other extreme, low contrast and dim
targets such as deep sky objects will be minimally affected because the fovea

cannot function at their low luminances. This leaves planetary observation and
close double stars widely different in magnitude as the targets where the size of
the central obstruction can be a critical issue, one reason that Newtonian or
refractor telescopes are often preferred by planetary and double star observers.

Optics of the Eye


The eye is an equal component with objective and eyepiece in the performance of
an optical system. The critical performance dimensions for astronomy are dark
adaptation, hue discrimination, and resolution as a function of stimulus brightness
and stimulus/background contrast. However I begin with a topic that is customarily
minimized or ignored in discussions of visual performance: individual differences.
Individual Differences
Human vision is generally characterized in three ways: by describing the physical
structure and theoretical optics of the eye; by calculating the optical performance
of "ideal" or mathematical eyes; and by measuring individual performance in a
variety of visual tasks.
Visual performance is measured using tasks of detection (the stimulus is declared
to be visible or not visible), matching (two stimuli are declared to be or are
adjusted by the observer until they appear to be the same), or discrimination
(similar stimuli can be identified as different in some way). In all cases, statistical
theory is required to summarize performance across individuals. Stimulus
thresholds in detection or discrimination are defined as the stimulus value that
produces a 50% or 95% probability that an average person will respond with the
necessary detection or discrimination. Matching stimuli are defined by the average
and standard error of stimulus values judged to be the same, or the average just
noticeable difference (JND) that produces detection that they are different. These
measurements are then used to create models of average human performance,
which deviate from actual performance.
Two examples will illustrate the scale of the problems. On the subject of star color,
the first diagram (below left) shows the location on a standard hue circle of the
range (variation) in the single wavelengths chosen as a "pure" or best shade of
red, yellow, green and blue light by subjects with normal color vision; the inset
shows the number of individual subjects (circles) who chose a specific wavelength
in the "green" matching task. On the subject of telescope resolution limits, the
second diagram (below right) shows the variation in the judged brightness of
wavelengths across the visible spectrum; the "peak" wavelength is used to define
telescopic resolution limits.

In the color matching task, there seems to be good agreement among subjects in
the locus of a "pure" yellow wavelength in fact, this constitutes the Nagal
anomaloscope test for red/green color blindness. But in the green and blue color
matching the performance can best be described as scattered: the green values
chosen by different subjects range from a "green blue" below 500 nm to a "green
yellow" near 560 nm.
In terms of photopic luminance, individual
peak sensitivity spans a range of at least
40 nm (from 535 to 575 nm). The
commonly used value of 555 nm is the
peak of the 1964 CIE V* photopic
luminosity function (orange curve) or a
similar function, which are actually not
defined by tests of brightness matching
but by a weighted average of the L and M
cone response curves measured with color
mixture matching.
These individual differences appear in all
visual tasks, however they are measured. Peter Kaiser and Robert Boynton
describe the situation this way: "Comparisons between different observers,
whether in the same or a different experiment, present a discouraging picture.
Although observers agree on certain major trends, individual differences are best
described as enormous."
Standard psychometric methods are designed to disguise what is typically an
"enormous" amount of individual variation in perceptual thresholds of detection
and discrimination. These individual differences vary with the type of detection or
discrimination required, on the specific way in which the visual stimulus is created
and presented, and most importantly, on the perceptual and cognitive attributes of
the observer, including past experience with the task it is possible to learn or
practice your way to better visual acuity.
If you are curious about your visual performance as a visual astronomer, it will be
very helpful to discuss testing for color normal vision, resolution and contrast
sensitivity with your ophthalmologist. It is easy to measure pupil aperture
yourself, when your eye is fully dark adapted. This information will provide an
empirical basis for your assessment of telescopic optical performance.
Illuminance & Luminance
For the purposes of describing a light stimulus, the key distinctions are between
illuminance and luminance.
Illuminance (E) is the quantity of light incident on a unit area of surface or
passing through a unit area of aperture, defined as:
E = LI/d2
where LI is the luminous intensity of the light (the quantity of light radiated into
a solid angle of 1 steradian, in lumens) and d is the distance in from the light
source. When distance is in meters and the unit area is a 1 meter square surface
or 113 cm diameter circular aperture, illuminance is measured in lux.

Luminance is the quantity of light emitted by a surface onto a surface, defined as:
L = E/s
where s is the solid angle of the light source (in steradians or square radians) as
observed from the receiving surface or aperture. When both the light source
surface and the receptor surface are standardized as 1 meter square, luminance is
measured in candelas per square meter.
In optical systems, the objective focal length o and aperture area D2 and
reciprocal relative aperture govern the relationship between light grasp and image
luminance:
Lo = D2E/o2 = E/N2
As explained above, changes in relative aperture produce squared changes in
image luminance.
The key distinctions between illuminance and luminance are:
Illuminance is completely invisible to the eye; all light perception is a form of
imaged luminance, perceived either as an emitting light source or a reflecting
material. All brightness and lightness perceptions are luminance perceptions.
Illuminance decreases as the light source is at a larger distance from the
receiving surface. Luminance decreases as the same quantity of light is emitted by
a physically larger source. Illuminance and luminance both decrease as the same
amount of light is incident on a larger surface area.
Illuminance provides no information about the size, distance or intensity of the
light source (the same quantity of illuminance can come from a nearby dim light or
a distant bright light). Luminance is a specific measure of the visual geometry of
the illuminance from a light source in relation to a light receptor; it depends both
on the angular area (surface area and distance) of the light source and on the area
of the light receptor (e.g., the diameter of the pupil), but is invariant with distance
because a light source that decreases in illuminace with distance decreases in the
same inverse square proportion in angular area.
Illuminance decreases with increasing distance between the light source and
receiving surface. Luminance remains constant regardless of the distance between
the emitting light and receptor surface the angular area of the source decreases
with distance in the same inverse square proportion that the illuminance from the
source decreases.
The power or light concentrating capability of an optical system is proportional to
D2/2 = N2; a 2" aperture astrograph operating at /4 delivers more concentrated
light to a photoreceptor than an /10 instrument of 20" aperture. Regardless of
aperture, the luminance of an optical image is equal to the luminance of the
source.
Illuminance and luminance are measurement units in photometry, which is limited
to electromagnetic radiation that produces a response in the eye, usually
considered to be wavelengths between 750 to 380 nm. Irradiance and radiance are
the equivalent measurement units in radiometry, which includes all
electromagnetic energy from radio waves through xrays. Thus, a star that radiates

very powerfully but almost entirely in either the infrared or ultraviolet will appear
visually faint, because little of the energy from the star can stimulate the eye.
Luminance Adaptation
Vision can be characterized by three adaptation states, each associated with a
range of environmental light levels (illuminance):
Scotopic. Only the rods respond to light, generally at illuminance levels below 0.1
lux. Several visual criteria define this adaptation, chief among them is the loss of
color sensation, although memory color serves to tint familiar objects near the
scotopic threshold.
Photopic. The cones predominate in the response to light, which nominally occurs
at illuminance levels above 10 lux or more. In fact, the rods are still fully
functioning at illuminance levels even 10 times higher, up to light levels typical of
noon sunlight. Photopic adaptation characterizes the state in which color
perception and visual acuity can be adequately described in terms of the cones
alone.
Mesopic. The transition adaptation between scotopic and photopic vision. Color
perception is present, but color saturation and contrast are muted, and hues are
somewhat biased by the rods.
Light adaptation is the response of the eye to increased luminance levels. It is
generally prompt, occuring within a fraction of a second at the pupil and within a
minute in the retina and brain.
Dark adaptation is the response of the eye to decreased luminance levels. occurs
within 20 minutes to 1 hour, depending on the light levels.
Light Grasp
The light grasp of the eye as a function of aperture is quite modest. At a dark
adapted pupil diameter of about 6 mm, the pupil area allows only 3.5% the light
grasp of a 150 mm (6") telescope. However the eye is capable, under optimal
conditions, of perceiving a light source emitting only a few photons per second.
Thus, what it lacks in aperture it makes up in retinal sensitivity.
Resolution
The minimum angular interval that can be resolved by a normal human eye varies
with the type of resolution required, but is usually given as Rv = 70 to 140
arcseconds (0.00034 to 0.00068 radians). This is approximately 6 to 3 pairs of
alternating black/white lines within 1 millimeter viewed from 25cm (152 to 76 line
pairs per inch, viewed from 10 inches).
The next diagram describes the eye's optical resolution in the abstract, as the spot
diameter of a point light source on the retina as a function of pupil diameter,
based on calculations with the Navarro ideal or mathematical "eye".

The Navarro eye allows


the spot diameter to be
separated into the
contributions from the
size of the diffraction
artifact (the Airy disc,
which gets smaller with
larger pupil aperture)
and from optical
aberrations in the cornea
and lens (geometric
spreading, which
increases with aperture
larger than ~3 mm). At
the pupil sizes
characteristic of most
visual astronomy, the
spot diameter in a
normal eye is generally
far in excess of 25 m (0.025 mm); but note that a pupil diameter of around
3.5 mm would apply to most lunar and some planetary observing. Values will
generally be worse for older and visually impaired observers.
A common method for estimating the eye's peak visual acuity is to calculate the
average spacing between two cones in the fovea as the diameter of a single cone,
and to assume that visual resolution is equal to this average spacing (the gap
between two point stimuli is as wide as one cone). The average cone density is
estimated to be around 190,000 per mm2 in the foveola (diameter on the retina of
about 0.35 mm), and about 100,000 per mm2 in the fovea (diameter of about
1.85 mm). This works out to an average cone diameter of about 0.0025 mm in the
foveola and 0.0034 mm in the fovea, given the characteristic hexagonal "tiling" of
cone aperture.
With the eye's average internal focal length of 24mm, these widths define an
angular separation of between 0.006 to 0.0081, or 0.36 to 0.49 arcminutes.
However, the minimum spot diameter of the eye under optimal conditions
(diagram, above) is about 7 m or 0.007 mm, which is roughly the diameter of
three cones. Thus the cone spacing method for predicting visual acuity overstates
the eye's detection threshold under optimal conditions by a factor of roughly 3
and this is consistent with the measured performance acuity of the normal eye
under bright light, which is between 1 and 2 arcminutes across the fovea.

However, the spot


diameter under low
luminance adaptation is
well above 0.025 mm,
which implies a visual
acuity in most
observational astronomy
that is as poor as 10
arcminutes. This
degradation of visual
acuity with increasing
pupil aperture is clearly
indicated in the
modulation transfer
functions of normal eyes
at various pupil apertures,
measured by
photographing the spot
size actually imaged on
the retina of subjects
(diagram, below).
Finally, the task dependence of visual acuity is reflected in the different
discrimination thresholds that characterize different monocular visual tasks:
Shape recognition - Black on white letter shapes can be resolved (e.g.,
rotational orientations of the shape "E" can be identified) when elements of the
letter are at least 5 arcminutes wide.
Grating resolution - Black on white line gratings (the lines and the spaces
between them are of equal width) can be discriminated from a uniform gray
background when the spacing between black lines is at least 2 arcminutes.
Point recognition - Two black points on a white background can be resolved
when the space between them is at least 1 arcminute; white points on a black
background can be discriminated at about 2 arcminutes (due to spreading of the
image).
Vernier (Nonius) acuity - The misalignment between two black lines placed
end to end or the vertical misalignment of two dots can be perceived when the
lateral displacement is at least 10 arcseconds. This form of acuity is the principle
behind the heliometer and the reason why it was the tool that first measured
stellar parallax (Bessel and 61 Cygni).
Contrast Sensitivity
Contrast or modulation range

The troland is a measure of retinal illuminance equal to the illumination from a


surface emitting 1 lux per square meter into a pupil area of one square millimeter;
it increases both with an increase in source brightness and an increase in pupil
aperture. For comparison, 1000 trolands is approximately the brightness of a
"white" paper viewed under bright indoor lighting, and 0.1 trolands is
approximately the brightness of white paper viewed under moonlight.
Most "deep sky" astronomical targets present retinal illuminances of less than 0.1
trolands, which implies a visual acuity limited at around 2 to 5 cycles per degree or
Rv = 360 to 900 arcseconds.
Field of View
Because the eye's normal point is the standard for visual magnification, calculation
shows that the eye's true field of view is equal to its apparent field of view.
As a simple test of your true field of view: stand and look straight ahead with both
eyes, then extend your arms on either side, fingers facing forward. Wiggle your
fingers while moving both your arms forward and backward until you find the point
where the movement of the fingers of both hands is just visible on both sides of
your visual field. The angle between your arms will be about 120.
However, the fact that your fingers are barely visible at 120 means that your
useable field is much smaller. How small? At the minimum, the fovea is only able
to resolve the central ~1 of your visual field. Perceptual research also shows that
viewers will say a single object fills their visual field when it subtends about a 20
visual angle a generalization first made by Leonardo da Vinci. For example, the
industry recommended wide screen TV viewing distances produce a visual
angle to the wide screen diagonal of between 20 to 40, with the average around
25 an image width that is said to produce "an immersive feel". By that
standard, the 40 apparent field of traditional eyepieces would be fully immersive.
Currently available commercial eyepieces offer apparent fields of view between
40 to 100. The choice of one end of that range over the other seems to tip at an
apparent field of around 60, and can depend on several considerations: esthetic
preference (the "majesty factor" as opposed to "looking through a drinking
straw"), visual task (sweeping for comets as opposed to resolving double stars),
the telescope mounting and balance sensitivity (super wide field eyepieces can be
very heavy, especially at long eyepiece focal lengths; super wide field eyepieces
allow extended "drift" viewing of objects with hand guided Dobsonian altaz

mountings), objective relative aperture (off axis eyepiece aberrations are more
noticeable at smaller focal ratios) and price (super wide field eyepieces are
generally much more expensive than traditional designs).
Off axis image quality, in telescopes with focal ratios down to about /6, can be
superior in some super wide angle eyepieces than in some traditional eyepieces,
and image quality at the center of the field (on axis) is, for almost all oculars
operating at all focal ratios, optically perfect. Only at objective focal ratios below
/5 do optical considerations become important.
On the other hand, super wide field eyepieces require the observer to "look
around" in the wide available field in order to see clearly, and this movement of
the eye and head can cause vignetting or aberrations to appear. (Bias
disclosure: I prefer eyepieces with apparent fields no greater than 70, mainly to
avoid these effects.)
In general, preference for wide or "narrow" apparent field eyepieces is not an issue
of optical quality or the "immersive" visual area useful to the steady gaze of the
human eye, but depends on the astronomer's personal preferences in visual
esthetics, ergonomics and visual tasks, the convenience of use with a specific type
of telescope ... and price.
Focus Diopters
Diopters are used to measure the distance that a lens must be moved forward or
backward along the optical axis to produce a change in the image focus equal to
one diopter of the eye. This focus diopter is found as:
e2/1000 mm.
Thus the movement required to adjust the focus of a 20mm eyepiece by one
diopter is (202)/1000, or 0.4mm.
Focus diopters may be positive or negative. At zero focus diopter, an eyepiece
projects collimated light rays of light parallel to the optical axis, which can then
focused by a relaxed normal eye as if the light came from a far distant source.
Therefore the optimal focus diopter for a normal eye is zero or very slightly
negative.

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic
optical principles.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of
historically important eyepiece designs.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces,
and results from my collection.
Amateur Astronomer's Handbook by J.B. Sidgwick - excellent basic chapters in
optics, light gathering, resolution, magnification and more.

Telescopic Limiting Magnitudes by Bradley Schaefer - an attempt to predict


telescopic limiting magnitudes using visual data, star color, age of observer, sky
brightness and other factors.
Human Eye from Handbook of Optical Systems, Vol. 4: Survey of Optical
Instruments by Herbert Gross, Fritz Blechinger & Bertram Achtner (eds.). (Berlin,
DR: Wiley-VCH, 2008).
Visual Acuity - Summary of the various methods used to test the human visus.
Average optical performance of the human eye as a function of age in a
normal population by A Guirao, C Gonzlez, M Redondo, E Geraghty, S Norrby
and P Artal. Investigative Ophthalmology & Visual Science Jan. 1999, pp. 203-213.
The N.A.A. Telescope Calculator - handy and accurate calculator page for most
optical parameters of a telescope/eyepiece combination.
Eyepiece Focal Length Measurement - Jim Easterbrook explains how to
measure eyepiece focal lengths in grating projections.
The Collimation - an excellent optical discussion by Thierry Legault
What Is a MTF Curve? - Basics of the MTF.

Last revised 6/20/12 2012 Bruce MacEvoy

Astronomical Optics
Part 3: Optical Aberrations

Chromatic Aberrations
Longitudinal Chromatic Aberration
Lateral Chromatic Aberration
Achromatic Optics
Monochromatic Aberrations
Defocus
Spherical Aberration
Coma
Distortion
Field Curvature
Astigmatism
Aberration Balancing
Spherical Aberration of the Exit Pupil

The Observer's Eye


Minimizing Optical Errors
Appendix 1: Graphical Representation of Aberrations
Wavefronts & Lenses
Spot Diagrams
Field Height Plots
Ray Intercept Curves
Appendix 2: Third Order Analysis
The Characteristic Function
The Seidel Sums
The Seidel Errors in Reflecting Telescopes

In their basic design, optical systems are held to the standard of first order or Gaussian optics: a
monochromatic "point" light source located at infinity and centered on the optical axis will appear as a
"point" image at the center of a focal plane that is flat and perpendicular to the optical axis. This standard is
then extended off axis to include the image of any point visible anywhere within the telescope image area or
eyepiece field of view.
Any departure from this optical perfection is an aberration. The most important of these were identified and
analyzed in the mid 19th century empirically by the Hungarian optician Joseph Petzval, and theoretically
by the German mathematician Philipp Ludwig von Seidel (pronounced ZYdul). They are usually called the
Seidel errors, and in optical systems that are symmetrical around an axis of rotation that is identical with
the optical axis, the Seidel errors are significant because they have both the greatest impact on image
quality and the greatest utility as guides to improving an optical design.
The five Seidel errors, in traditional order, are: (1) spherical
aberration, (2) coma, (3) astigmatism, (4) field curvature and
(5) distortion. Two types of first order (6,7) chromatic
aberration (caused when the image is not monochromatic) are
consistently included among the important aberrations; and (8)
spherical aberration of the exit pupil is a flaw often
encountered in wide angle eyepieces.
Aberrations can be analyzed and classified in several ways. As
shown in the chart (left), a simple distinction is between
aberrations that fail to produce a point image for every object
point in a common focal plane (termed errors of focus), or
aberrations that fail to produce an equal image scale either at
the image point or across the entire image area (termed
errors of magnification).

in short focal ratio objectives, long focal


eyepieces (60 or greater apparent field of

Astigmatism combines features of focus and magnification


error and is intimately associated with both field curvature
and distortion. All three tend to become more pronounced
length (low magnification) eyepieces and wide field
view).

Aberrations may appear in light rays


entering the lens close to and parallel to the optical axis
(termed paraxial rays); or they may appear only in light rays entering the system near the edge of the
objective or at an angle to the optical axis (off axis or abaxial rays). The abaxial errors become more
significant as the distance between the object image and the center of the field of view (of the objective
and/or the eyepiece) increases.

The starting assumptions are that all the refracting or reflecting surfaces are either flat or are spherical
surfaces of rotation around an optical axis, the aperture and field stops are circular and perpendicular to
the optical axis, and all stops and surfaces are centered on the optical axis.
These assumptions mean that the system is perfectly collimated. Collimation is the procedure of aligning
the objective optics with the eyepiece optics, focal reducer or camera focal plane so that they share a single
optical axis.
Apart from atmospheric turbulence, aberrations in astronomical optics may appear at any of three points
the telescope optics (objective lens, or primary and secondary mirrors), the eyepiece and related optics
(such as a field flattener, focal extender or focal reducer), and the observer's eye. Severe misalignment
between the observer's eye and the instrument optical axis are common, and can produce aberrations in
the center of the field. Practical experience indicates that the eyepiece or photographic lens generally
dominates the contribution of aberrations in telescope systems, and that thermal turbulence in the
instrument and atmosphere limit the magnification and visual or photographic resolution to "what the
seeing will allow."
Optical Parameters & Units
Discussion of aberrations depends on a few technical parameters and terminology, summarized in the
diagram (below) and reproduced from the page Telescope & Eyepiece Combined. See there for further
explanation.

Aberrations are calculated Each of the eight aberrations is illustrated below with a diagram that simulates its
"pure" appearance in a telescope field of view. These are not photographs of actual aberrations and are
intended only to aid visual recognition. They do not characterize the aberrations as they appear in all
situations; typically, two or more aberrations will affect an image at the same time and will appear in
different proportions at different field heights within the image.
Chromatic Aberrations
Chromatic aberration is a failure to focus or magnify the image equally across all spectral wavelengths or
spectral hues ("colors"). It arises because optical materials slow the long (red) wavelengths of light into a
less extreme angle of refraction than the short (violet) wavelengths, visibly spreading out the spectral hues
in an optical effect called dispersion.
Chromatic aberrations are considered the most objectionable (visually most confusing or distracting)
aberrations in an optical system, and designs to minimize chromatic errors appeared very early. Pursuit of
chromatically better corrected optical designs was a principal reason for the development of new types of
optical glasses in the 19th century.

Chromatic aberration appears in two forms. Longitudinal chromatic aberration (above, left) is an error of
focus; it occurs when different wavelengths of light from a single object point are refracted into separate
focal planes at different focal lengths along the optical axis. Lateral chromatic aberration (above, right) is an
error of magnification, which occurs because different wavelengths of light from an off axis object point are
dispersed at different powers from opposite sides of the aperture.
Longitudinal Chromatic Aberration
Longitudinal or "axial" chromatic aberration focuses different spectral hues at different distances along the
optical axis, with "violet" wavelengths focused at a point closer to the lens than "red" wavelengths. The
separation between these extreme spectral focal points defines the secondary spectrum of the focused light.
This spectral separation causes the "best focus" bright point images to appear tinged with green and
surrounded by red fringes, and edges between light and dark areas to be fringed with red or violet color.

A ray tracing (geometrical) analysis of longitudinal chromatic aberration (diagram, above) shows that the
refracted light does not define a single focal plane. Instead the focal point defined by yellow, which creates
the visually brightest image, is usually chosen as the best focus.
At the "yellow" focus both the red and violet wavelengths are located outside the apparent edge of an
object image, where they mix to create a red violet fringe, while the blue, green, yellow and orange
wavelengths mix within the object image to produce (for a white object) a white color tinged with green.
From this best focus, the intrafocal image presents only red at the outside fringe and produces a central
image tinged with blue, while the extrafocal image presents an outside fringe of blue violet and a central
image tinged with red.
As an axial optical error, longitudinal chromatic aberration affects all parts of the image roughly equally.

Lateral Chromatic Aberration


Lateral or transverse chromatic aberration appears in off axis image areas. It occurs in images brought into
perfect refractive focus because the lateral spectrum produced by dispersion is of different widths or
magnifications in light arriving from opposite sides of the aperture. This produces radial color within the
meridional plane, particularly for highly contrasted points and edges, as shown in exaggerated scale in the
diagram (below).

Because the spectra are overlapped in reverse order, mixing complementary hues, this light mixture mostly
produces a "white" focal image. However light from the opposite side of the aperture arrives at a longer
focal length and is therefore dispersed at a slightly higher magnification, so that it extends beyond the focal
"white" image on both the "red" and "violet" ends. Typically the "violet" end is magnified to a greater extent
so that it appears larger than the "red" end nearer the optical axis.
The spread in dispersion is always a small fraction of the refractive angle, so lateral color in well corrected
achromat optics is usually at a scale similar to the Airy disk diameter and is only visible at high
magnifications and sharp edges.
Lateral color is affected by the location of stops or silhouetted objects such as etched or wire reticules. In
optics where the aperture stop (in objectives) or field stop (in eyepieces) is in front of the lenses, the "blue"
fringe is located on the opposite side of the image from the center of the field; these are termed
overcorrected. In systems where the aperture or field stop is inside or behind the lenses, the "blue" end of
the spectrum is located on the same side of the image as the center of the field; these are termed
undercorrected.
The entire image area will show lateral color at the image of the edge of the field stop; in overcorrected
systems the edge of the field will be tinged with blue, or tinged with red in undercorrected systems. (The
color near to the optical axis does not appear because it overlaps with dispersed light from the inner parts
of the image field.) In optimally corrected systems the field stop will appear fringed with yellow green.
As an abaxial optical error, lateral chromatic aberration increases as the angular height () or field height
(h) increases.
Achromatic Optics
The method for minimizing chromatic aberration in refractive optics is to use two lenses whose combined
refraction and dispersion powers are manipulated to "fold" the secondary spectrum back onto itself, so that
the extreme ends of the spectrum focus in the same focal plane.
Seventeenth and 18th century refractors were built to extremely slow focal ratios (/100 and longer)
primarily to minimize axial color. The Huygens eyepiece, a combination of two lenses that minimizes
lateral color through the spacing of two positive lenses, was developed late in the 17th century.
The strategy to eliminate lateral color by means of a doublet composed of a positive and negative lens was
developed independently in 18th century England by Chester Moore Hall and John Dolland for eyepiece

optics, and brought to a high refinement in large aperture telescope objectives by Joseph von Fraunhofer
and Carl Steinheil. These doublet lenses are termed achromatic.
Dollond's son Peter conceived the addition of a third lens so that the middle of the achromatic spectrum is
folded on itself again, resulting in an improved achromatic system, termed apochromatic (diagram, below).

Note that in relative terms, the reductions in chromatic error are roughly equivalent about 80%
compression of the secondary spectrum at each step. But in absolute terms (the amount of defocus along
the optical axis) there is a gross improvement from an uncorrected to achromatic system, but a much less
dramatic improvement from achromatic to apochromatic.
In naked eye experience, fringing is a strong signal of luminance intensity, as dispersions appear more
colorful ("prismatic") in bright lights. But when the orange/blue contrast fringes light areas against a dark
background, it has a strong spatial or depth effect on the image that is distracting and confusing (see
image, above).
Both forms of chromatic aberration are absent in nondiffracting (smooth mirror) reflective optics.
Monochromatic Aberrations
After the chromatic aberrations are the five most serious monochromatic aberrations, all errors of refraction
rather than dispersion. These affect the image focus or magnification even when the image is
monochromatic transmitted in a single wavelength of light. First in importance among these are spherical
aberration and coma.
Defocus, Decenter & Tilt
Three forms of optical misalignment affect the first order or Gaussian image point.
Defocus is the longitudinal displacement of the effective focal plane from the first order or Gaussian focal
point. This causes a radial expansion of the image point on the effective image plane, decreasing the
luminance of each point and causing point images to overlap.
Decenter is the displacement of an optical element so that its center is no longer on the optical axis.
Tilt is the rotation of an optical element around an axis that passes through the center of the element and
is perpendicular to the optical axis.

Collimation is the process of adjustment that eliminates tilt and decenter from the
optical system.
Spherical Aberration
Spherical aberration (diagram left) is a failure of focus, caused when rays parallel
to the optical axis have different focal lengths at different aperture heights s. It
produces a roughly uniform, concentric defocus around all the bright elements in
all parts of an image, and a blurring in extended and low contrast detail,
preventing any part from coming into sharp focus. The aberration is uniform
across the entire image area.
The analysis of spherical aberration is related to its appearance in the spherical
reflection of collimated rays (diagram, below). Each aperture height s defines a
concentric area of the objective with a specific (and different) focal length.
Undercorrection or negative spherical aberration occurs when rays incident at a large aperture height
(reflected from the near edge of the objective) have a shorter focal length than the central rays; this is
characteristic of spherical mirrors and converging lenses.
Overcorrection or positive spherical aberration occurs when the peripheral rays have a longer focal length
than the central rays, which is chatacteristic of ellipsoidal mirrors and spherical diverging lenses.

The terms probably arise from the traditional method of figuring a Newtonian mirror, which is first ground
out to an approximately spherical figure with the desired focal length, then polished or "corrected" into a
paraboloid figure by flattening the outer surfaces. However the terms apply equally to lenses or mirrors of
any conic shape.

Because the light cones of the rays cross through each other as they come in and out of focus, spherical
aberration produces some defocus at every point along the optical axis. The best available focus, in which
the rays that appear brightest to the eye (hues cyan to orange) are contained within the smallest area, is
usually (in undercorrection) extrafocal from the circle of least confusion, where the diameter of all the

intersecting light cones is at a minimum.


As the diagram (above) shows, at an extrafocal position (with the eyepiece focal
plane closer to the observer than the least circle of confusion), the core of the
image displays more of the peripheral rays; at an intrafocal position (the least
circle of confusion closer to the observer than the eyepiece focal plane) it displays
more of the central rays.
In many instances, spherical aberration produces a "best focus" or "least
confused" image of a star that is only marginally larger than the Airy disk
produced by diffraction: most of the out of focus light appears as a brightening of
the rings around the disk. For this reason, spherical error does not significantly
degrade the resolution power of a telescope in double stars of similar magnitude,
but brightening of the rings does impair resolution for binaries of unequal
magnitude, as the rings of the brighter star will mask the fainter. Spherical aberration also significantly
impairs the resolution of low contrast detail, for example in planetary and deep sky objects.
Coma
Coma (diagram, right) is a failure to magnify equally off axis light rays passing through the objective or
eyepiece. Unlike spherical aberration, the effect is not constant within the same concentric aperture height
height (s); instead the magnification also varies with field height (h). It produces elongated, wedge
shaped "tails" that extend radially from point images or luminance edges in off axis light.
The aberration occurs because light from opposite sides of the objective are reflected or refracted at
different powers, due to the different angles of incidence at the optical surface. The severity of the
aberration increases with field height (h) of the object point, with the more distorted rays being incident
at a higher aperture height (s) of the objective. Coma can also appear in on axis object images if the
optical system is not correctly collimated.
Coma is inherent in parabolic mirrors, because parallel but off axis rays of light do not strike opposite
sides of the mirror at equal angles of reflection: on one side of the mirror the angle become decreases it on
the opposite side; in effect, it is as if one side of the mirror becomes flatter, and the other side more curved
(diagram below, left). A similar effect occurs as an imbalanced refractive power on opposite sides of a lens,
due to the different angles of incidence of parallel but abaxial rays entering the lens on opposite sides
(diagram below, right).
In external coma, the more common form, this produces a greater magnification of star images as the field
height of the incident rays increases, causing a series of images to appear both larger in diameter and at a
greater distance from the center of the field: the tails of these comatic blurs point toward the field edge. In
internal coma, the tails point toward the field center.
As shown in the illustration (above), coma also becomes more extreme as the field height increases: point
images near the edge of the field are more strongly blurred than images near the center, and the comatic
blurring increases with the larger angular heights contained in low power images.

The apex of the coma is defined by the rays passing closest to the principal point of the lens (red lines,

above). The radial length of the coma, aligned to the axis of the aberration, constitutes the tangential
aberration, and the diameter of the widest part the sagittal aberration. Their lengths are related in the ratio
3:2, and are proportional to the angular height of the point image times the square of one half the focal
ratio: tan()(D/2o)2. Most of the aberrated image is contained within the brightest "V" rim of the figure,
expanding as a 60 wedge from the point formed by rays passing through the center of the objective.
In modern optics, coma principally results when the eyepiece and/or objective are not perfectly collimated
(their optical axes are exactly aligned), or when the focal ratio is fast (the focal length is a small multiple of
the aperture, typically below /5).

After the chromatic aberrations, spherical aberration and coma were the optical errors next successfully
tackled in eyepiece designs. To identify this achievement, the 19th century German optician Ernst Abbe
introduced the term aplanatic to describe eyepieces free of both aberrations. Aplanatic optics obey the
constant e = h'/sine() for all collimated rays, known as the sine condition (diagram, above).
This is equivalent to the condition that the light rays are refracted to the focal point along paths of equal
length, as if perpendicular to the surface of a spherical "principal plane" (shown as a circular arc in the
diagram) whose radius is the focal length (e).
With the sine condition satisfied, distortion, astigmatism and field curvature remained the aberrations
constraining the width of field (maximum angular height ) in astronomical optics.
Distortion
Distortion is caused by a failure to magnify the image equally as the field height increases.

Distortion is characterized in two ways. In positive or rectilinear distortion, the magnification increases
toward the edge of the field: this causes straight lines to appear curved outward, even when the image is
stationary. In negative or angular magnification distortion, the magnification decreases toward the edge of

the field: this causes straight lines to appear projected onto a spherical surface curved toward the viewer,
which is especially noticeable when the optical axis is moved the image then appears to "roll away" near
the field edge.

The sign of the weight assigned to distortion can be remembered by visualizing the effect on a square
placed with one corner on the optical axis, so that the two sides and diagonal originating from the corner
are radial lines from the center of the image. Since the diagonal is longer than a side, it has a greater field
height and will show a greater alteration by the distortion.
If the distortion coefficient is positive (rectilinear distortion), then the diagonal is made longer and the far
corner of the square becomes pointed (less than 90); hence it is sometimes called pincushion distortion.
If the coefficient is negative, the diagonal is made shorter and the far corner becomes larger or rounded
(greater than 90); this is sometimes called barrel distortion.
The two types of distortion cannot be eliminated at the same time: given an ocular field height (h') and
angular height at the exit pupil (), both in radians, then rectilinear distortion will be zero when the image

is scaled as h' = etan(); but angular magnification distortion will be zero when the image is scaled as h'
= e(). Visually, rectilinear distortion is the effect produced when a plane image is projected onto a
spherical surface, and angular magnification distortion is the effect when a spherical image is projected onto
a plane surface.
Geometrically, making them equal amounts to making both the side and diagonal of a square the same
length. Obviously, then, either positive or negative distortion must be allowed if it is necessary to reduce
the other.
Some amount of positive (rectilinear) distortion, which causes straight lines to appear bent toward the edge
of the field, is usually introduced to minimize or eliminate the negative (angular magnification) distortion,
which causes circular objects (such as planets) to appear squashed or stretched out near the edge of field.
A common marketing tout for wide field eyepieces is: "Our eyepieces are absolutely free of angular
magnification distortion!" Unfortunately, this is achieved by significantly increasing the rectilinear distortion.
In addition, increased astigmatism is often the result of minimizing distortion in pursuit of a "flat" field: the
magnification error is treated as separate sagittal and tangential components, which allows them to be
manipulated separately.
In 19th century and early 20th century eyepiece designs, especially for military sighting optics, elimination
of rectilinear distortion was critical for use of the optics to measure field angles as indicators of distance. In
astronomical optics, elimination of angular magnification distortion is usually critical, as the "roll off" of the
image as the telescope is moved is considered unesthetic and distracting.
Field Curvature
Field curvature (called Petzval curvature in the technical literature) is a failure to focus the entire image on
a single plane perpendicular to the optical axis. Instead the focal "plane" is a paraboloid surface resembling
a bowl or meniscus. This produces a characteristic inability to focus the center and edges of the field at the
same time.

Field curvature is directly related to astigmatism, as the average of the sagittal and tangential curvatures
(see below, and Appendix 2). Almost all astronomical objectives and eyepieces present some combination
of curvature of field, astigmatism or distortion, especially wide field optics.
The curved principal focal plane is termed the Petzval surface, and the amount of curvature is indicated by
the best fitting spherical radius of the curvature, the Petzval number. A smaller Petzval number indicates a
shorter radius and therefore a greater degree of curvature in the focal plane.
Nearly all telescope objectives produce an image with positive curvature, in which the edges of the field are
focused at a shorter focal length than the center of the field (the bowl is turned toward the objective and
away from the observer). However, in eyepieces, where the focus is at the image plane, positive curvature
means the bowl is facing toward the observer. Thus, if the objective has a positive curvature (as most
telescope objectives do), and the eyepiece has a negative curvature, then the two focal surfaces will curve
in the same direction. significantly near the edge of the field, especially in low power eyepieces or low focal
ratio objectives.

Positive curvature in the objective image causes objects at the edge of the field to lie at an intrafocal
position in relation to a focused field center in the eyepiece focal surface. In an SCT, once the field center is
brought into focus, the focusing knob must be turned clockwise to bring the edges into focus. In negative
curvature, the edges of the field are focused farther from the objective than the center (the focal bowl faces
the viewer), and the edges are brought into focus by a counterclockwise rotation of the focusing knob.
A classic Schmidt Cassegrain camera produces a large negative curvature of field, which is compensated by
using a photosensitive surface with a matching convex surface.
The eyepiece tangential and sagittal astigmatic curves can be manipulated to counteract some field
curvature, usually at the cost of introducing astigmatism. Studies of eyepiece/objective combinations by
Rutten & Van Venrooij (1988) indicate that eyepiece astigmatism is responsible for most of the astigmatism
in any objective/eyepiece combination, and that eyepiece astigmatism overwhelms the coma produced by a
Newtonian reflector, particularly toward the edge of the field.
Astigmatism
Finally, astigmatism occurs when light rays from perpendicular cross sections of the image cone do not have
the same focal distance along the optical axis. It is therefore an error both of focus and of magnification.
Astigmatism is pervasive in optical systems, and occurs in all abaxial light passed through any refracting
lens. It is the most difficult aberration to correct.

Astigmatism produces two focal points for every abaxial image point. These are described in terms of two
planes governed by the position and field height of the off axis object (diagram, below).
The first plane contains both the optical axis and the principal ray from the off axis object, and divides the
focal cone in half along the optical axis. This is the meridional plane, and the fan of rays within it defines the
tangential focus.

The second plane also contains the principal ray of the image point but is perpendicular to the tangential
plane; it only intersects the optical axis at the principal point. The fan of rays within this plane defines the
sagittal focus.

Geometrically, the difference between the two components of astigmatism is easiest to visualize as separate
diagrams for an object located directly below the optical axis, in the direction indicated by the white arrow.
In this diagram, the path of the sagittal rays appears in the view from above (yellow lines) and the
tangential rays in the view from the side (green lines). Note that within each plane, the opposing plane is
perpendicular and appears as a single line because it is viewed edge on.
Viewed from above, the sagittal plane is symmetrical around the optical axis. Although the sagittal rays
strike the lens at equal angles to the surface normal on opposite sides of the lens, the abaxial tilt in relation
to the optical axis effectively thickens the lens: the rays must travel a longer distance within the lens before
reaching the opposite side, and therefore meet the lens at a point where its curvature in relation to the
optical axis is less extreme and its refractive power is therefore reduced. This causes the sagittal rays focus
long as a very thin line or ellipse perpendicular to the optical axis, which contains the aberration in the
tangential rays (diagram above, left).
Viewed from the side, the tangential plane is symmetrical around the principal ray of the object. As a result,
the tangential rays strike the lens at different angles to the surface normal on opposite sides of the lens,
changing its effective figure and producing different focal lengths in the opposing rays in a manner similar
to coma. This causes the rays to focus short as a very thin line or ellipse radial to the optical axis, which
contains all the aberration in the sagittal rays (diagram above, right).
The "best focus" is located approximately midway between the sagittal and tangential focal points, and
produces an approximately circular and defocused image of a point source which contains the tangential
and sagittal aberrations in equal proportions.
The complexity of astigmatism is that the sagittal aberration is governed by a symmetrical change in focal
length, while the tangential astigmatism is governed by an asymmetrical difference in focal length.
Astigmatism therefore combines errors in focus and magnification: the sagittal and tangential images come
into focus at unequal distances from the optical axis, so they are at different powers (magnifications).
The two components of astigmatism change in focal length, and typically diverge from each other, as as a
function of the square of the angular height of the image (2). As a result they define two separate, bowl
shaped (rather than flat) focal surfaces, that exactly coincide or intersect at the optical axis. In field height
diagrams they are represented as a radius cross section through each bowl, which produces two separate
curves for the sagittal (S) and tangential (T) surfaces (diagrams, below).

In these diagrams, the vertical scale is the field height of the focused rays and the horizontal scale is the
difference in focus diopter adjustment from the paraxial focus.
Note that either or both forms of astigmatism can define a positive or negative curvature, and that either
type of astigmatism can be more extreme (produce a larger defocus) than the other. In the geometrical
diagram (above), the tangential rays focus in front of, and the sagittal rays focus behind, the focal point
defined by paraxial rays; but the tangential rays can also focus behind the sagittal rays, and both can focus
in front of or behind the focal point of paraxial rays.

The field curvature plots (above) show the amount of tangential and sagittal astigmatism in the illustrative
lens designs, where the horizontal scale is diopters of focus, the vertical scale is field angle, and the
objective is on the left. This illustrates that the tangential astigmatism is much more variable than the
sagittal: in uncorrected systems, the tangential curve is always farther from the Petzval surface than the
sagittal curve.
The visual effects of astigmatism can be reduced by manipulating the sagittal and tangential surfaces to
counterbalance each other. In the simplest strategy to do this, the tangential surface is made positive and
the sagittal surface negative (or vice versa), so that their average is close to zero at all field heights (middle
two diagrams, above). In another strategy, the outer "rim" of either the sagittal or tangential astigmatic
"bowl" is turned backward so that passes back through the curvature of the opposing dimension, producing

zero astigmatism at some constant radius (usually around 70% of the maximum field height) from the
optical center and minimizing the divergence between the two dimensions as much as possible. An eyepiece
with minimal or no astigmatism is called anastigmatic.
When the concentric smearing of sagittal astigmatism is combined with coma, it creates the effect
photographers call sagittal coma flare, which induces a "flying bird" appearance to star images at increasing
field height, shown as both ray spot and the visually more accurate point spread functions in the diagram
(below).

Seidel astigmatism is not the same as the common astigmatism that affects the human eye, which occurs
even for paraxial light and is caused by the cornea and/or lens having different curvatures in different cross
sections.
Aberration Balancing

In most modern eyepiece designs, distortion is primarily a difficulty in wide field or ultra wide field
eyepieces, with maximum angular heights () of 30 or greater. In eyepieces with apparent maximum
angular heights of 20 or less, the difference between an angle and its tangent is visually so small (as a
difference in h') that distortion does not become noticeable (chart, below).

Systems that have minimized both angular magnification and rectilinear distortion are termed orthoscopic:
specifically they satisfy the magnification constant Me = h'/tan(). When distortion occurs, the orthoscopic
formula alters to:
Me = h'/[tan()(1+Etan()3)]
where E is the coefficient of distortion. This formula shows that distortion increases as the cube of the
angular height (3), as shown in the chart (above).

aberration

vs. aperture

vs. field angle

spherical (longitudinal)

D2

spherical (lateral)

coma

field curvature (longitudinal)

field curvature (transverse)

astigmatism (tangential)

astigmatism (sagittal)

distortion (percent)

longitudinal chromatic

lateral chromatic

Since the tangent that defines distortion and the sine that defines coma produce different ratios, an
eyepiece cannot be both orthoscopic and aplanatic. However, as already mentioned, the differences among
the ratios defined by the field height (h'), orthoscopic (tangent ) and aplanatic (sine ) quantities diverge
noticeably for values of greater than about 20 (chart, above).

Spherical Aberration of the Exit Pupil


Seidel errors can apply to the image formed within the eyepiece just as in the objective, and these
contribute to the aberrations already in the objective image. However, spherical aberration presents itself
visually not as an aberration of focus but as an aberration of position, known as spherical aberration of the
exit pupil or "kidney bean blackout" of the image field.
This situation arises because the eyepiece is overcorrected: it focuses the peripheral rays at an image plane
in front of the central rays, as shown below.

The eye is optimally positioned with the exit pupil centered within the pupillary aperture, which forms the
entrance pupil of the eye (upper diagram). So long as the eye is fixated at the center of the field, and the
pupil is placed at the focal plane of the central and near central rays (red and yellow lines in the diagram),
the peripheral rays (blue lines in the diagram) fall on the sides of the retina where imaging resolution is
very poor, and the lack of focus is not noticeable. However, if the eye is moved forward and rotated to
admit the peripheral rays from one side of the image field in order to view that side of the field distinctly
(lower diagram), the peripheral and near central rays from the opposite side are blocked by the
foreshortened pupil or the wall of the eye, producing a mobile, kidney shaped "blackout" on that side of the
visual field.
Spherical aberration of the exit pupil only makes obvious a more basic constraint: the portion of the total
visual field of the eye that is capable of forming a distinct image is relatively small. Extreme detail can only
be imaged at the fovea, which subtends about 2 at the fixation point, and then only for light that is above
the mesopic threshold (roughly, above 0.1 candelas per square meter). The parafoveal area, about 20
wide, is the image width commonly preferred by observers who are asked to adjust an extended object
image to the largest size that still lets them view the object as a whole, when fixated at its center. This
implies a physical diameter to distance ratio of about 1:3 equivalent to viewing an audio compact disc
(12cm diameter) at a distance of 36cm from the eye.
The smallest apparent field offered today in commercial astronomical eyepieces is about 40, twice the
parafoveal image area and a diameter/distance ratio of 1:1.4 (a CD viewed from 16 cm). Many types of
eyepieces offer apparent fields extended up to 60 (approximately 1 radian, or 57.3), a diameter/distance
ratio of 1:1.15. At this extreme width, the parafoveal area can be directed to one side or the other of the
available field without overlapping the central image area, and when the eye is fixated to one side of the

image area the opposite side becomes completely indistinct.


It's worth mention here that there are at least three distinct forms of "shadowing" or partial obscuration of
the eyepiece field (diagram, below).

Spherical aberration of the exit pupil produces a crescent or oblate shadow, mobile within the field, that
appears both detached from the edge of the field and surrounded at its edges by smeared out star images.
A similar type of shadow, appearing as a crescent edge that occludes the field from one side, occurs
because a fully flat exit pupil is not completely encircled by the eye pupil (the eye's optical axis is displaced
to one side, or the eye is turned sharply to one side). Finally, if the pupil is smaller than the central shadow
of the secondary mirror within the exit pupil, this appears as a circular shadow in the center of the field,
which can entirely black out the field in extreme cases. This is most common when using a low
magnification ocular under daylight illumination, which causes the pupil to contract to a diameter of about
one or two millimeters.
Spherical aberration of the exit pupil almost exclusively afflicts eyepiece designs with fields larger than 60
wide angle or super wide angle eyepieces. Given the limitations of the eye, and the fact that spherical
aberration will black out a large part of the eyepiece field, rendering it useless, it is more practical to use
eyepieces that present the true field of view as an apparent field no wider than 60, then choose an
eyepiece focal length (magnification) that fits an extended object entirely within the eyepiece's true field of
view.
The Observer's Eye
For many observers, the most significant source of optical aberrations is not the astronomical objective or
eyepiece but their own eye(s), so these deserve mention as a source of optical errors. Unfortunately no
conclusive description can be given for eyes in general or for an individual eye specifically, because the eye
adjusts to the visual stimulus by changing the shape of the lens (accommodation), changing the diameter of
the pupil, and/or changing the response sensitivity of the photoreceptor cells and neural networks in the
retina (adaptation).
As an added complication, the resolution of the eye is not a single fixed value because it depends on the
specific type of discrimination being measured. For example, discriminating different shapes, such as block
letters, can be done down to about 5 arcminutes; discriminating a grating of alternating sinusoidal
black/white lines from a flat gray area can be achieved down to about 2 arcminutes; separating two closely
spaced points can be done down to 1 arcminute; recognizing a sideways misalignment between the ends of
two parallel straight lines (Vernier discrimination) can be done down to 10 arcseconds; and recognizing a
spatial separation between the edges of two objects at different distances can be achieved down to a
binocular difference of 5 arcseconds.
As a benchmark, we can assume that the resolution limit for line spacings at mesopic levels of luminance,
viewed within the 2 to 4 visual area of the fovea around the fixation point, is approximately 1 arcminute,
or 1/3438th (0.0003) of a radian. This quickly declines to 0.002 radian (~7 arcminutes) at a field angle of
10 and to 0.005 radians (~17 arcminutes) at a field angle of 30. Performance becomes worse as the
luminance falls to scotopic (starlight) levels, increasing the effects of abaxial aberrations such as
astigmatism and chromatic aberration, which contribute 1 to 3 diopters of defocus at a 30 field angle in a
normal eye. This is in addition to the eye's negative field curvature, which conforms to the spherical surface
of the retina. (Astigmatic eyes will contribute more defocus, and nearer the fovea.)

***check*** Defined as a point spread function, the average eye reaches its peak optical quality at a pupil
diameter of about 3.5mm, where the Airy disk projected on the retina is about 7.6 micrometers wide. At
this peak, the Airy disk stimulates around 10 foveal cones (spaced at around 2.3 cones per micrometer).
Below this diameter, the Airy disk enlarges quickly, to about 0.000022 mm at 1mm (~90 cones); above this
diameter, the geometrical spreading of the spot due to abaxial aberrations expands it to 0.000028 mm at
6mm diameter (~150 cones). Note that the telescope exit pupil serves to stop down the aperture of the eye
in exactly the same way as the iris.
The eye suffers from significant longitudinal chromatic aberration, and the secondary spectrum has a spread
of about 2.5 focus diopters: from yellow (peak sensitivity) red is focused at +0.5 diopters and violet at 2
diopters. This is suppressed by the blue filtering effects of the yellowed lens and maculate pigmentation,
and by neural processing. Lateral chromatic aberration is also present, especially when the exit pupil of a
telescope and the pupil of the eye are not aligned exactly, but the added chromatic aberration is only
around 0.04% per millimeter of displacement.
Minimizing the Optical Errors
These eight optical defects are related mathematically (see Appendix, below) in ways that make it
impractical if not impossible to eliminate them all in an astronomical eyepiece or objective, especially when
the goal is a wide field design or the focal ratio (/D) is small. It is therefore necessary to balance one
aberration against the others in optical designs in order to produce the best optical performance.
Because the abaxial errors lateral chromatic aberration, coma, astigmatism, distortion, curvature of field
and spherical aberration of the exit pupil become more pronounced as the field height (h) or angular
height () of the light source increases, the most distorted part of any astronomical image is toward the
periphery of the image area, and the best corrected part of the image is at center of the image, around the
optical axis.
The two chromatic aberrations, spherical aberration and coma are generally considered the most
objectionable image defects, and all are consistently almost entirely removed in modern telescope designs.
In systems where the defects are unavoidable astigmatism in a Ritchey-Chrtien reflector or coma in a
Newtonian or Dall-Kirkham reflector the field of view is limited (due to a large relative aperture) to the
area around the optical axis where the aberrations are so small that they are below the visual or
photographic angle of resolution.
Field Size & Collimation
Aberrations are a function of the focal ratio, the angular width of the apparent field of view (in the
eyepiece), and the magnification, or angular width of the true field of view.
Since Galileo's time, optical defects in a telescope have commonly been minimized by stopping down the
telescope aperture, a practice used by early 20th century visual astronomers such as Percival Lowell to
reduce the effects of atmospheric turbulence. In photographic cameras with very short focal ratios, even
those with 12 or more spherical and aspheric lenses, stopping down is the function of the diaphragm or field
stop, which limits light to the area around the optical axis of the system, minimizing off axis aberrations.
Even in these systems, abaxial aberrations can still be evident in point source images located toward the
edge of the field of view and photographed in dark environments when the diaphragm must be fully opened.
Wide field eyepieces with long focal length (low magnification) create the most serious demands on an
optical system and are most likely to show some combination of coma, astigmatism, distortion and field
curvature at field widths greater than 60.
In the same way, objectives with a fast relative aperture are especially susceptible to aberrations,
particularly coma, astigmatism and distortion.
Aberrations can also be produced by optical systems that are out of collimation or alignment to a single
shared optical axis.
Collimation includes the viewer's eye, and aberrations can emerge when the optical axis of the eye does not
coincide with the optical axis of the exit pupil. These appear in the centered disk of a moderately bright
star, and also in the form of a slightly defocused star image.
A third source of aberrations is atmospheric turbulence, which can produce a dizzying sequence of
defocused, comatic, astigmatic, and displaced (distorted) aberrations in a perfectly focused star image.
Under poor seeing these follow one another so quickly that the image degrades into a boiling ball of light,
fraught with rapidly shifting and dissolving, very thin and very crisp diffraction shadows.
Focus and Resolution
Nearly all modern optical systems introduce some form of optical aberration into their images. The key issue
is whether these are noticeable or bothersome if they have any impact on the intended use of the
telescope.

Photographic imaging generally requires more stringent performance than visual systems. The eye's retinal
resolution is roughly 2 arcminutes, and taking the Airy disk diameter as a standard unit of angular error
within an optical system, aberrations in the objective become noticeable at roughly 2 diameters in
photography and 3 diameters in naked eye observing.
Defocus is the most primitive or basic form of optical aberration, and precise focus is critical to good optical
performance. In many systems, defocus also projects underlying optical errors more clearly, so that a star
image just out of focus may appear comatic or astigmatic. In reflector telescopes, the shadow of the
secondary mirror produces a "donut" shaped ring of light in a defocused star image. The inner and outer
borders of this figure will be chromatically tinged with a slight residue of lateral chromatism. These fringes
reverse position inside and outside of best focus: in undercorrected systems, the extrafocal image will
appear red around the interior and blue around the outside, reversing position in the intrafocal image.
Appendix 1: Graphical Representation of Aberrations
It may be useful to review the mathematical analysis of the Seidel (third order monochromatic) aberrations,
as this can clarify how they are produced by an optical surface.
Wavefronts & Lenses
Eyepiece lenses are constructed as solids of rotation, which means their refracting surfaces are defined
geometrically around an axis of rotation. Most are spherical surfaces, although aspheric lenses use surfaces
such as ellipsoids, paraboloids or hyperboloids (respectively an ellipse, parabola or hyperbola rotated
around its major axis).

In the thin lens model the radius of the Petzval curvature () measured from the focal point is equal to:
= n'
A positive lens produces a surface concave to the left (diagram, above), and a negative lens produces a
surface concave to the right.

Spot Diagrams
The spot diagram is commonly used in optical design to assess image quality. It is created by
tracing a large number (usually around 100) of rays from a point source that are incident in a
regular or geometric pattern across the entire aperture.
The spot diagram plots the location of these rays where they intersect the image plane. The
diagram (below) shows the spot diagrams for "pure" forms of spherical aberration, coma and
astigmatism, in focal planes that bracket the optimal focal position.

Usually several aberrations affect the spot at once. For that reason, the spot at best focus is
assessed in relation to a criterion spot diameter, which indicates the maximum spot dimension
that is acceptable for a specific application. In visual optics the criterion can be as small as the
diameter of the Airy disk (equivalent to the Rayleigh 1/4 criterion); for photographic
applications the criterion can be larger, on the order of 0.025 mm.
The point spread function is the wave diffraction equivalent of the spot diagram. An example
point spread function is illustrated here. More complex to calculate, it is used to provide the
most accurate and detailed quantitative assessment of optical quality.
Field Height Plots
Only two meridonial paraxial rays are necessary for the analysis: the axial ray which intersects
the aperture at the same field height as the point source, and the principal ray through the
principal point at the aperture stop (diagram, below).

Ray Intercept Curves


Finally, aberration diagrams can also be based on the trigonometry of meridional rays. These are
known as a ray intercept curve or H'tan U' curve.
The diagram (left) shows the basic logic of this kind of graph. Each meridional ray is plotted as
the field height h' of its exit from the last refracting surface and the tangent of its angle u' to a
reference plane that is perpendicular to the optical axis and placed some arbitrary small distance
behind the optimal focal point.
The inset diagram shows that a perfectly formed image will have

Ray intercept curves also allow interpretation of the effects of refocusing the image, shown as
the slope of the line or curve relative to the horizontal (x) axis. Refocusing rotates the curve
around its central point, either counterclockwise (for intrafocal adjustment) or clockwise (for
extrafocal adjustment).
In the example (diagram at right, top), the curve for the undercorrected zonal spherical
aberration is presented, with dashed lines showing the amount of aberration at the optimal focus
(that is, the focus that most reduces the aberration in the central part of the image), and dotted
lines showing the amount of aberration at the focus that reduces the total aberration.

The second diagram (left, bottom) shows the effect of refocusing the image intrafocally to the
optimal criterion the dashed line is now parallel to the x axis. Most of the rays are refocused to
within the criterion error. Note however that the amount of defocus in the total light is larger,
since the slopes for the optimal and total refocusing in the upper diagram are not equal.
The ray intercept curves show that coma cannot be reduced or eliminated by refocusing: focus
only determines which part of the image will be most out of focus. Refocusing can significantly
reduce spherical aberration, and either tangential or sagittal astigmatism (at the expense of the
other). Refocusing cannot eliminate axial or lateral chromatic aberration, although it does
determine which part of the spectrum will appear in focus.
Appendix 2: Third Order Analysis
It may be useful to outline the mathematical analysis of the Seidel (third order monochromatic)
aberrations, as this can clarify how they are produced and interrelated.
Only two paraxial rays from an object point O are necessary for the analysis: the principal ray
through the principal point at the aperture stop, and a marginal ray that intersects the aperture
space at point A. Both rays converge after refraction to the image point I (diagram, below).

The principal ray defines the meridional plane, which always includes the object point O, the
principal point P, the Gaussian image point I, and the optical axis. Not shown in the diagram is
the sagittal plane, which is perpendicular to the meridional plane and includes the principal ray
(points O, P and I) instead of the optical axis.
Because all lenses are surfaces of rotation around the optical axis, the direction in which the
object point is off axis doesn't matter only its off axis radial distance h has any importance. So
for convenience the analysis framework is arranged so that the y axis in object space and the y'
axis in image space lie in the meridional plane: y = h.
The aperture space uses a different convention for the incidence point A: its location is defined
by s, the distance from the principal point, and , the angle measured at P between A and the
meridional plane. This angle is expressed in radians: for rays in the meridional plane = 0 or pi;
for rays in the sagittal plane = 1/2pi or 3/2pi.
All points A that are not within the meridional plane are skew rays, and can represent any light
ray emanating from O that is incident on the entire surface of the aperture. For astronomical
objects (at optical infinity) the skew rays are parallel to the principal ray and both the meridional
and sagittal planes.

The significance of using a principal and a marginal ray is that aberrations are created primarily
by light rays at a distance from the optical axis: either rays that are incident near the margin of
the lens, and those incident near the margin of the image.
Now consider the implication of various values of h, s and :
The significance of h is that it determines the angular height at the principal point, which
describes the abaxial angle of the column of light originating at h. For the object distances and
true fields of view typical in astronomical work, h can be considered identical with either angle
expressed in radians or the tangent of . Since angle is equal on both sides of the lens, h also
determines the field height of the point in the image plane.
The off axis distance h shifts the angle of incidence of all rays originating at O in the same
direction and by an equal amount (the angle ). In the sagittal plane, the angles of incidence for
matching rays on opposite sides of the principal ray remain equal. But because the meridional
plane contains the optical axis, rays above the optical axis are inclined away from the axis while
rays below the axis are inclined toward it, so that the angles of incidence for matching rays on
opposite sides of the refracting surface can be quite unequal.
The significance of s is that it defines the distance of the incidence point A from the vertex of
the refracting surface at P. This indicates greater curvature in the lens surface: as the marginal
ray and principal ray diverge by a larger distance (and s increases), the refractive power of the
lens at point A must also increase in order to make the rays converge again at the image point I.
This refraction is characterized as a zone that is concentric around the principal point at a
constant radius s. For all equal values of s the refracting power of the optics will be equal, and as
the value of s increases the concentric refracting power also increases.
The significance of is that its sine and cosine represent the relative proportion in the
refractive effect at A that can be attributed to the sagittal tilt or the meridional tilt induced by h.
Thus, cos() is 1 or 1 when the ray is in the meridional plane and 0 when the ray is in the
sagittal plane; and sin() is 1 or 1 when the ray is in the sagittal plane and 0 when it is in the
meridional plane. (Mnemonic: sine is sagittal.) In other words, the sine and cosine of define
the X,Y coordinate locations of all points on a circle that is concentric around P and has a unit
radius s.
The Characteristic Function
An intuitive analytic formulation of optical aberrations was developed by William Rowan Hamilton
in 1833 and is known as the characteristic function. All light rays from the object point,
regardless of their incidence point in the aperture area, must focus at a single point in the image
plane, located by the principal ray on the y' axis at a height proportional to h. Any nonzero value
of x' is by definition an aberration; any value of y' that is not proportional to h is also an
aberration.
The characteristic function identifies the origin and form of the departure from the Gaussian point
focus x' and y' as the weighted sum of different combinations of the normalized field height of
the object point (h, where the field limit equals 1.0), the normalized aperture height of any
incident ray (s, where the aperture stop equals 1.0), and the aperture azimuth angle of any
incident ray (, where the meridional plane is = 0 or 180). The parameters h and define
the type of aberration; s shows whether and how the effect varies over the aperture area.
In this form the Seidel errors are technically known as third order aberrations because their
mathematical definition requires terms for field height (h) and aperture height (s) whose
exponents sum to three h3, hs2, h2s or s3. (Fifth order and seventh order aberrations are
sometimes analyzed in complex or asymmetrical optical systems.)

Relative
Orientation

Gaussian

Spherical
Aberration

Coma

Astigmatism

Curvature

Distortion

y'

meridional

A1scos()+A2h

+B1s3cos()

+B2s2h(2+cos(2))

+3B3sh2cos()

+B4sh2cos()

+B5h3

x'

sagittal

A1ssin()

+B1s3sin()

+B2s2hsin(2)

+B3sh2sin()

+B4sh2sin()

The size or magnitude of each aberration is indicated by the subscripted weights A1, A2 and B1 to
B5, and these weighted aberration terms add together to indicate the total aberration in the point
image.
Because the principal ray passes through the origin of the aperture coordinate system, s equals
zero. All the aberration terms drop out by multiplication with 0, and only the Gaussian value A2
dependent on h remains: the field height of the image point I is linearly proportional to the
angular height of the object point O. Any deviation from this point by the marginal ray image
point, and the terms that contribute to this deviation, define the type and degree of aberration.
Gaussian. The Gaussian parameters for x' and y' in A1 do not contain h: therefore there is no
noticeable difference in image focus or image quality between on axis (h = 0) and off axis (h >
0) image points. The refractive power is linearly related to s, and there is no difference in effect
between the meridional and sagittal planes. These are the characteristics of Gaussian or "perfect"
optics.
The radial term B5 dependent on h<sup3< sup=""> does not drop out for the principal ray and is
independent of the aperture incidence point A of any marginal ray. This reflects the nonlinear
increase in magnification for objects closer than 2 to the aperture plane. </sup3<>
The Gaussian sagittal (x') and tangential (y') terms are nonzero primarily in cases of defocus or
longitudinal chromatic aberration. In those cases the size of the aberration is dependent only on
the refractive power s, and produces a circular enlargement of the image that is concentric
around the principal image point.
Spherical aberration. This is the only third order aberration independent of the object angular
height h: it can therefore appear even in on axis image points and affects all image points
equally. Thus, the aberrated rays are symmetrically distributed around the image point (the
parameter forms are identical in the x' and y' equations).
The aberration increases in proportion to the cube of the aperture height (s3). If we assume that
the aperture area is equally illuminated, then the exponent means that most of the visible
aberration is generated by the areas with large s that are near the marginal circumference of the
aperture, and this marginal effect is much more pronounced for s3 than for s2.
That interpretation always assumes that the focal plane is located at the Gaussian image point.
As illustrated above, moving the focal plane along the optical axis will bring the rays from
different aperture heights into focus, so that for example the visible aberration can be made to
arise from the central aperture area at the marginal ray focal point. In other words, refocusing
will change the size and even algebraic sign of the aberration weights A and B.

Coma. The complexity of the


coma aberration is evident
from its parameters in the
characteristic equation. We
see first that coma is related
to both h and s<sip2<
sup="">: coma becomes
linearly larger as field angle
h increases, and increases
with the refracting power s2
toward the marginal
circumference of the optics.
Thus coma appears in off
axis image points, and
proportionately most of the
visible aberration arises from the marginal rays. </sip2<>
However, the aperture orientation is doubled for both the x' and y' dimensions, which means
that the comatic image results from the superposition of light rays from opposite sides of the
aperture area and concentric around the principal point P, as shown in the diagram (right).
In addition, the tangential contribution is shifted by a factor of 2, regardless of the value of or
s, which means the center of the circular aberration is shifted radially away from the principal ray
image point I by a distance equal to twice its diameter. (Because always has a unit diameter,
the coordinate locations of B and D will be x' = 1 and 1, A and C will be located at y' = 3 and
1.) The last two factors produce the distinctive shape of the comatic aberration, which consists of
circular focal rings nested within a 60 angle expanding from the principal ray image point. The
aberration is brightest at the point and along the sides, producing a "webbed V" shape.
The exponential terms s2 and s3 indicate that spherical aberration and coma are much more
strongly affected by the shape of the refractive surface than by the off axis angle of the incident
rays. That means both aberrations are always present in single lenses with spherical surfaces.
The aberrations must be corrected by making a single lens or reflecting surface aspheric either
paraboloid or hyperboloid or by the use of compound lenses with different radii whose
combined effect approximates an aspheric power.
Astigmatism. Like coma, astigmatism is a complex aberration with asymmetrical contributions to
the x' and y' dimensions. The tangential weight is 3 times the size of the sagittal, which means
they focus at different locations; and these are in relation to the focal surface defined by the
Petzval curvature the astigmatic tangential focal surface is 3 times farther from the Petzval
surface than the sagittal focal surface.
Petzval curvature. The curvature parameter is identical in form to the astigmatic, dependent on
both aperture zone s and the field angle h2. Unlike astigmatism, the x' and y' contributions are
identical. Petzval curvature defines a paraboloid focal surface whose vertex is at the principal
focal point that becomes more extreme with the field angle.
Distortion. This aberration is a tangential (radial) displacement of the image as field height
increases; it has no sagittal component. It increases as the cube of the field angle.
The Seidel Errors in Reflecting Telescopes
Listed below are the calculations for the five Seidel aberrations, as derived from basic
calculations and as given by Nicklas (1994). These reveal the dependence between the Abbe
invariant (Q) and spherical aberration (I), spherical aberration and coma (II), and finally the
dependences among astigmatism (III), field curvature (IV) and distortion (V), which are all
three affected by the Petzval sum () and the position of the aperture stop (p).
Each surface of the optical system contributes an additive portion of the total aberration, and is
denoted by a subscript: h1 denotes the first surface, and h denotes each of the subsequent
surfaces.
Notation
n refractive index of air or glass
h intersect height of refracted ray on lens surface
s distance from surface intersect to optical axis measured, from surface vertex along optical
axis

u angle of refracted ray to optical axis


r radius of curvature of refracting surface
d focal length of lens
dEP distance from surface vertex to aperture stop
x signifies a quantity before refraction (to the left of refracting surface)
x' signifies a quantity after refraction (to the right of refracting surface)
Preliminary Quantities
Dispersion
(1/ns) = 1/(n's') 1/(ns)
Refraction

h/h1 = (s2s3s4 .. s)/(s'1s'2s'3 .. s'{1}) =

Aperture Stop Position

Abbe Invariant

Q = n(1/r 1/s) = n'(1/r' 1/s')


p = 1/((h/h1)2Q) +
= (1/r)(1/n' 1/n)

k=2..

j=2..(sj

/ s'{j1})

[ dk / (nk(h{k1}/h1)(hk/h1)) ] dEP

Petzval Sum
Seidel Partial Coefficients
Spherical Aberration
I = (h/h1)4Q2(1/ns)
Coma
II = pI
Meridional astigmatism
IIIa = 3p2I+
Sagittal astigmatism
IIIb = p2I+
Astigmatic Difference
IIIc = p2I
Mean Field Curvature
IV = 2p2I+
Distortion
V = p3I+p

= 3IIIc+
= IIIc+
= (IIIaIIIb)/2
= (IIIa+IIIb)/2
= p(IIIc+)

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 2: Telescope & Eyepiece Combined - the optics of astronomical telescopes.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of historically important
eyepiece designs.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces, and results from my
collection.
Intrinsic Telescope Aberrations - intro page to Vladimir Sacek's discussion of the Seidel errors and
Zernicke analysis.
Starizona's Telescope Basics - A straightforward ray tracing explanation of optical aberrations.
What Is Aberration? - Visual and geometric illustrations of the Seidel and chromatic errors.
The Light Fantastic: A Modern Introduction to Classical and Quantum Optics by I.R. Kenyon.
Primer of Image Aberrations and Their Representation - Tutorial on the graphical analysis of
aberrations.
"Aberrations" in J.B. Sidgwick, Amateur Astronomer's Handbook. (Dover, 1955).
"Optical Telescopes and Instrumentation" by H. Nicklas. In Compendium of Practical Astronomy, Vol. 1
edited by Gunter Rth (Springer-Verlag, 1994). Overview of astronomical optics with specific sections on
the aberrations.
Handbook of Optics: Design, Fabrication and Testing by Michael Bass, Virendra N. Mahajan & Eric Van
Stryland. Detailed and very clear explanation of optical aberrations and their graphical representation.
Fundamental Optical Design by Michael J. Kidger. Excellent basic chapter on aberrations.
Optical Imaging and Aberrations by Virendra N. Mahajan. Theoretical and mathematical, but with much
specific practical information.

Last revised 5/15/12 2012 Bruce MacEvoy

Astronomical Optics
Part 4: Eyepiece Designs

Design Considerations
Field Lens & Eye Lens
Apparent Field Categories
Pre 19th Century Designs
Kepler
Huygens
Hall
Ramsden
19th Century Designs
Barlow/Smyth Lens
Kellner
Plssl
Monocentric
Orthoscopic

20th Century Designs


Hastings Triplet
Erfle
Kaspereit
Knig
Brandon
Zeiss Astroplanokular
RKE
Panoptic
Takahashi LE
Nagler
Pentax XW
Ethos

A "Gleanings for the ATM" column in the February, 1944 Sky & Telescope began by observing
that "eyepieces are commercially available, but not in any particular profusion of types or focal
lengths, and not always easily procured." How times change! The amateur astronomer is likely to
be staggered by the variety of eyepiece brands, designs and price points available today. While
there might be three basic types of telescopes, there are dozens of generic designs,
proprietary designs, and manufacturing variations of eyepieces.
Design Considerations
The ideal eyepiece is succinctly defined as crisp, wide, flat, bright, dark, comfortable, durable and
affordable. This means the eyepiece must produce an image with minimal or no aberrations
within the range of objective focal ratios it is designed to magnify; it must provide a wide field
without perceptible curvature, transmit almost all light from the image without ghosts, glare or
scatter, provide enough eye relief for comfortable viewing (preferrably with an adjustable eye
rest), withstand years of normal use, environmental exposure and the occasional accident
without affecting any of the previous qualities, and be offered new at a reasonable price.
The first five constitute the optical performance of the eyepiece, the first seven constitute the
manufacturing quality of the eyepiece, and the last constitutes the market availability of the
eyepiece. The market availability is affected by business and marketing costs, market demand
and manufactured supply and is irrelevant to design issues. The optical performance is partly
dependent on the optical design but also and importantly on the manufacturing tolerances and
the quality of raw materials (optical glass) used in fabrication, and the focal length specification
of the eyepiece. The optical performance of an eyepiece is a product of the optical design and of
the skill and care with which the design has been made into a physical object required to magnify
at a specific focal length.
The development of the modern eyepiece has been a history of progress on three fronts: the
optical theory necessary to optimize the crisp, wide and flat attributes of the image; the optical
materials and coatings necessary to optimize the optics with a bright and dark presentation; and
the fabrication technology necessary to create the object at reasonable cost. Quite often
fabrication issues or materials costs require compromise in the optical design.
Due to a variety of proprietary restrictions and marketing misnomers, the details of eyepiece
optics are difficult to penetrate with assurance. In nearly all cases I was unable to find
prescription data for the earliest example of an eyepiece design, and nearly all the designs have
been tweaked significantly since they were invented through the use of newer optical glasses and
small changes in the optical proportions. I have tried to select design schematics that use a 25
mm focal length.
The following summary is based principally on the chapter "Eyepieces" the Handbook of Optical

Systems, Volume 4: Survey of Optical Instruments edited by Herbert Gross, Fritz Blechinger &
Bertram Achtner; the chapter "Eyepieces for Telescopes" in Harrie Rutten & Martin van Venrooij's
Telescope Optics; the chapter "Oculars" the Handbook of Optical Systems, Volume 4: Survey of
Optical Instruments edited by Herbert Gross, Fritz Blechinger & Bertram Achtner; and "The
Evolution of Eyepiece Design" by Christopher Lord, along with many other primary and secondary
sources.
Adopting Lord's eyepiece format, each illustration gives the eyepiece name and year of design or
patent, with the design apparent field of view, eye relief as a proportion of the focal length (e),
and the fastest objective focal ratio.
Field Lens & Eye Lens
The diagram (below) illustrates the two basic strategies of eyepiece design. In eyepieces
constructed on the 19th century or standard design, the eyepiece can be divided into two
components which serve two distinct functions. The field lens is weakly positive, and
concentrates the peripheral (abaxial) light rays (blue lines) so that they pass through the eye
lens. The eye lens is strongly positive, and defines the apparent field of the eyepiece. This type of
design is not used for apparent fields greater than about 50.

In contrast, many wide field designs can be divided into three functional units. The field lens is a
negative Smyth group (effectively, a built in Barlow lens) that is actually brought in front of the
telescope focal plane, sometimes followed by a negative lens after the field stop. This diverges
the peripheral rays to an even higher field angle, where they are first approximately converged
by the central elements of the eyepiece and then brought into the exit pupil by the eye lens at a
steeper angle, thereby forming a wider apparent field (typically 70 or more).
Note that astronomical oculars are sometimes referred to as inverting eyepieces. In fact, as a
species of magnifier, eyepieces produce images that are both erect and normal when used by
themselves: the "inverted" (actually rotated) image is created by the objective, not the eyepiece.
This usage apparently originated in the 17th century with Schryleus, as a contrast term to
"erecting eyepiece" (also known as a terrestrial eyepiece or image erector).
Apparent Field Categories
Eyepieces can be classified by the width of their apparent field of view, or the eyepiece field as a
virtual window. This is separate from the true field of view, which is the eyepiece field as an area
of the sky. The apparent field is generally divided into three categories: (1) traditional or
standard eyepieces have an apparent field of view from as little as 25 up to about 1 radian
(57); (2) wide angle eyepieces have an apparent field of view from 60 up to 80; and (3)
super wide angle or ultra wide angle eyepieces have an apparent field above 80 (the largest
field achievable is around 120).
The following table lists the apparent field widths that can be simulated by viewing a standard
compact disk (12 cm in diameter) at the given distances.

Pre 19th Century Designs


The earliest telescopes were empirical constructions produced without a theory of optics or an analytical approach to telescope
design. Lenses were ground and put together in various combinations to find those that worked. Galileo wrote: "Spyglasses that
are most exquisite and capable of showing all the observations [described in his Siderius nuncius of 1610] are very rare, and
among the sixty that I have made, at great cost and effort, I have been able to find only a very small number." The quality of optical
glass and lens manufacture was poor, so lenses could be neither large nor thick; telescopes were made with focal lengths of 100
feet or more, and routinely stopped down in aperture, to minimize the negative effects of chromatic aberration and poorly figured
optics. The story of pre 19th century eyepieces traces the efforts necessary to master the three basic challenges of optical theory,
optical aberrations and lens manufacture.
Kepler. The biconvex lens is sometimes cited as the earliest form
of magnifier. In fact, the ocular in the astronomical telescopes
constructed by Gailileo Galilei from 1609 to 1621 was a plano
concave lens, following the designs manufactured by the Dutch
and Parisians. The biconvex ocular is first described in the Dioptrice
(1611), a discussion of telescope optics by Johannes Kepler (15711630), which was published to affirm and also clarify the physical
basis for Galileo's observations. Though he reasoned with only an
approximate theory of refraction, Kepler showed that a telescope
made with a biconvex objective and ocular produces a magnified
image when the focal points of both lenses coincide. This has the
potential for a larger apparent and actual field of view than was
possible with Galileo's telescope, and by extending the radius of curvature for one side of the lens it is
possible to reduce somewhat both chromatic and spherical aberration but unlike the Dutch design it
produces an inverted image. The field of view beyond 15 was badly distorted by off axis aberrations; eye
relief is approximately the eye lens focal length.
Huygens. The brilliant Dutch mathematician Christiaan Huygens
(1629-1695, pronounced Hoyghenz) was author of Trait de la
Lumire (1690), which summarized optical theories he began
developing in the 1650's and was the first optical treatise to apply
the law of refraction to lenses of spherical surface and the design of
telescopes. In 1662 Huygens developed the eyepiece that bears his
name: it consists of two plano convex crown lenses with both
curved surfaces facing the objective, mounted with a spacing
between the lenses equal to half the sum of their separate focal
lengths, which minimizes chromatic aberration, and with the two
focal lengths in the ratio 3:1 (field:eye), which minimizes spherical
aberration; however two common early designs utilized the ratios
3:2 (for high power magnification) and 4:1 (for low power). The Huygens is a negative eyepiece, meaning
that it cannot be used as a simple magnifier (to examine a postage stamp or insect, for example). It
places the objective image plane inside the eyepiece (between the two lenses) where it is transmitted to
the eye with the uncorrected chromatic aberration of the eye lens; as a result, use of a reticule or
crosshairs becomes impractical as these will be blurred and fringed with color. The huygenian eyepiece
has significant spherical aberration, field curvature and some negative (pin cushion) distortion and coma;
it also has slight negative astigmatism, which can be used to counteract the negative astigmatism of a
high focal ratio (> /12) objective. It works best with refracting telescopes. The design originally had a
25 to 30 apparent field of view and very short eye relief less than 8mm at e = 28mm. English
astronomer George Airy minimized the spherical aberration and field curvature by using a positive
meniscus field lens and a biconvex eye lens; German optician Moritz Mittenzwey widened the field to 50
by using a positive meniscus field lens and a plano convex eye lens. Despite its antique origin, the
huygenian eyepiece is still sometimes used in professional high focal ratio refractors, which minimize its
optical flaws.

Hall. John Dollond (1706-1761) was an English manufacturer of


fine navigational and "philosophical" (scientific) instruments. After
years of experimentation, he developed an achromatic doublet
which he described to the Royal Society in 1758 and patented for
manufacture a year later as both achromat objectives and
eyepieces. It consists of a biconvex (positive) lens of lower index
crown glass cemented into a plano concave (negative) lens made
with higher index flint glass. The two lenses were designed so that
their relative dispersions counterbalanced to eliminate chromatic
aberration while their refractive indexes combined to focus the
image. Used as an ocular, the doublet has a 20 field of view and
eye relief of about 26mm at e = 28mm.
Ramsden. Jesse Ramsden (1735-1800) was Dollond's son in law,
learned from him the manufacture of optical and precision
instruments, and founded his own instrument manufacturing
company. His eyepiece design consists of two plano convex crown
lenses of equal focal lengths, separated by about 2/3d the sum of
their focal lengths and with the plane surfaces facing outward
(away from each other). The Ramsden is a positive eyepiece that
can be used as a simple magnifier, meaning that the focal plane is
in front of the field lens. In this form the Ramsden design has a 25
apparent field of view and better correction than the Huygens for
spherical and off axis aberrations, though with some residual lateral
chromatic aberration. But it also has zero eye relief and all flaws of
the field lens (surface scratches and dirt, enclosed bubbles) appear in focus. To mitigate these drawbacks,
the spacing and focal lengths of the lenses are modified away from the optical ideal, which yields an eye
relief of about 7mm (at e = 28mm) but introduces significant field curvature, ghosting and lateral
chromatic aberration (color correction can be improved by the choice of glasses); however the external
focal plane does allow this eyepiece to be used with a reticle in long focal ratio telescopes.
19th Century Designs
At this stage the problems of chromatic and spherical aberration were well appreciated and increasingly minimized in optical
instruments, and the technology of precision machine manufacture was capable of producing scientific instruments of unparalleled
excellence. Microscopy and daguerrotype photography expanded the range of optical requirements and applications, and these
often laid the foundation for telescope eyepiece design. Nineteenth century optical designers of eyepieces were concerned with
increasing the field of view and eye relief, shortening the focal length, and reducing further the optical errors that persisted in 18th
century designs. These efforts were advanced in the mid 1800's through techniques of mathematical optical design and aberration
analysis developed by Joseph Petzval (1807-1891) and Philipp Ludwig von Seidel (1821-1896). Perhaps most important, after
1830 a greater variety and quality of optical glasses were available for experimentation and combination from manufactories such
as Guinand (France) and Chance Bros. (England); innovation accelerated again after 1886. These gave optical designers greater
control over refraction and dispersion and new avenues for innovation. German entrepreneurs founded some of the first major
optical manufacturing companies (Zeiss, Leitz, Steinheil), and the profit motive meeting end user requirements and minimizing
manufacturing costs shaped lens designs as well.
Barlow/Smyth Lens. The idea of using a negative doublet to
extend the virtual focal length of an objective or flatten a curved
field was introduced more than once in the 19th century. Peter
Barlow (1776-1862), an English mathematician and engineer,
developed the negative achromat with George Dollond, who
presented it to the Royal Society in 1834. Charles Piazzi Smyth,
Astronomer Royal for Scotland, conceived the idea of using a
negative achromat to minimize field curvature in an otherwise
aberration corrected wide field lens. After World War II, the barlow
lens became a standard tool for multiplying from 1.5 to 3 times the
magnification of eyepieces (for
objectives with focal ratios longer than f/6); the Smyth lens was
adapted to correct wide angle field curvature in 20th century lens designs the 110 military binocular
eyepiece developed by Tronnier in 1943, the 110 periscope eyepiece patented by Khler in 1959, the
Pretoria eyepiece designed by Klee & McDowell to compensate for aberrations in an f/4 parabolic reflector,
the astonishingly well corrected "extreme wide field" (90) eyepiece patented by Don Dilworth in 1988,
and the several commercial designs by Al Nagler and Explore Scientific.

Kellner. Carl Kellner (1829-1855) was a German mathematician


and machinist. The Kellner is basically a Ramsden modified by
replacing the plano convex eye lens with an achromatic doublet,
which effectively eliminates chromatic aberration. It is probably the
oldest eyepiece design still used in binoculars and marketed to
amateur astronomers; the various aplanatic eyepieces, which
correct both spherical aberration and coma, were developed from
it. It has very little longitudinal chromatic aberration and very low
astigmatism, field curvature and distortion; its spherical aberration
can be minimized by modern optical glasses, and its tendency to
excessive ghosting can be controlled with lens coatings. It offers a
very sharp, bright central field at low to medium powers, but very
short eye relief at high powers (short focal lengths). Sometimes a biconvex lens is substituted for the
plano convex field lens. The Kellner design typically has a 40 to 45 apparent field of view but with eye
relief of 13mm (at e = 28mm).
Plssl. Georg Simon Plssl (1794-1868, pronounced Plursul) was
an Austrian scientific instrument maker. His design is derived from
the Dialsight or symmetrical eyepiece, which consists of two
identical achromatic doublets, oriented so that the biconvex crown
elements face each other and spaced at roughly 20% of their focal
lengths. (The kinship of this concept to a triplet magnifier appears
if we imagine joining the two central crowns as a single crown
biconvex lens.) However, the symmetrical is prone to ghosting, a
problem Plssl addressed by bringing the doublets almost into
contact, increasing the refraction of the field lens flint element, and
reducing the diameter of the field doublet; these modifications were
further developed and patented by Knig in 1939. In these versions
the negative flint elements are slightly concave on the exterior surfaces. Through alternative choice of
glasses, these can be made flat, and all curved surfaces of equal radius, but with noticeable loss in
performance. The design shown has modest distortion of about 8%; lateral color, spherical aberration and
coma are completely corrected, and ghosts are almost entirely absent. The largest residual aberration is
field curvature, which can be corrected somewhat by increasing the eyepiece astigmatism. Plssl
eyepieces typically offer a superior field of view to the orthoscopic: variants can reach a 50 apparent field
of view and eye relief of about 20mm (at e = 28mm).
Monocentric. Hugo Adolph Steinheil (1832-1893) was son of Carl
August, inventor of the first miniature camera, and from 1855 head
of the family's optics company. His monocentric design is an
elaboration of spherical eyepieces, made by dropping beads of
molten glass into hot water, that were used by microscopist
Anthony van Leeuwenhoek and astronomer William Herschel. These
led to a variety of spherical or rodlike single element lenses devised
in the 18th and early 19th centuries by Wollaston, Brewster,
Coddington, Stanhope, Tolles and others; all were limited by very
short eye relief and usually also by spherical and/or chromatic
aberration. Steinheil perfected the concept by fitting two concentric
flint caps over a crown core; all surfaces are spherical around a
common center. Steinheil's version is almost completely achromatic with very slight spherical aberration
and a flat, very dark field free of ghosts; it has a narrow 25 to 30 apparent field of view, best at f/6 and
above, and substantial eye relief of about 24mm (at e = 28mm). Sidgwick describes the monocentrics as
"without doubt the most nearly perfect oculars that have yet been designed".
Orthoscopic. Ernst Abbe (1840-1905, pronounced Abbay and in
England sometimes incorrectly spelled "Abb") was a brilliant
German mathematician and physicist. He introduced the
orthoscopic design in 1860. It consists of a triplet field lens with a
single biconvex or plano convex eye lens. It shows a longitudinal
spherical aberration of about 0.5 diopters, no coma, sagittal
astigmatism of about 1.2 diopters and very little tangential
astigmatism. It was the first eyepiece with almost completely
corrected and distortion (less than 4% at a field height of 8), hence
the name (Greek for "straight seeing"). It offers excellent
sharpness, color correction, and contrast, and improves on the
Kellner design with unusually good eye relief. This design has a 45

apparent field of view and eye relief of about 22mm (at e = 28mm).
20th Century Designs
Eyepieces of the 20th century were strongly influenced by military requirements, where manufacturing
cost is no consideration and a distortion free, very wide field is essential. The flood of surplus eyepieces
that appeared after every major war stimulated commercial interests to imitate and innovate from them,
and several defense contractors (Brandon, Scidmore, Knig) patented unique designs. Late in the century,
computer assisted design and ray tracing programs put much more complex solutions within reach, but as
a result ultra wide field eyepieces became almost absurdly large and heavy. The most recent innovations
suggest that minimizing the mass of glass, the number of lenses and the number of air/glass or
glass/glass boundaries necessary to produce a wide field has become a priority, at the same time that
computerized lens manufacture in Europe and Asia has greatly driven down the cost of exotically figured
(aspherical) surfaces.
Hastings Triplet. Patented for Zeiss in 1911, this is a modification
by Paul Rudolph (1858-1935) and Charles Hastings (1848-1932) of
the symmetrical and monocentric triplet devised by Steinheil in
1860. It is today one of the most common forms of the loupe or
pocket magnifier, but despite that humble occupation it is close to
diffraction limited and is the basic design of the TMB Monocentric.

Erfle. Heinrich Erfle (1884-1923, pronounced Airfluh) was a


German physicist who worked for Steinheil & Shne and Carl Zeiss.
His design marks the transition to 20th century and the first truly
"wide field" eyepiece, which was not coincidentally created for
military applications (and was widely available as war surplus in the
1950's). Patented in 1923, the three erfle eyepieces consist of 5
elements in a 1-2-2 or 2-1-2 arrangement that offers long eye
relief but a relatively close alignment of field lens to field stop. The
distortion (for similar angular fields) is comparable to that of the
orthoscopic. This design has a 60 apparent field of view and eye
relief of about 20mm (at e = 28mm).

20th Century Designs

Eyepieces of the 20th century were strongly influenced by military requirements, where manufacturing cost is no consideration and
a distortion free, very wide field is essential. The flood of surplus eyepieces that appeared after every major war stimulated
commercial interests to imitate and innovate from them, and several defense contractors (Brandon, Scidmore, Knig) patented
unique designs. Late in the century, computer assisted design and ray tracing programs put much more complex solutions within
reach, but as a result ultra wide field eyepieces became almost absurdly large and heavy. The most recent innovations suggest that
minimizing the mass of glass, the number of lenses and the number of air/glass or glass/glass boundaries necessary to produce a
wide field has become a priority, at the same time that computerized lens manufacture in Europe and Asia has greatly driven down
the cost of exotically figured (aspherical) surfaces.

Hastings Triplet. Patented for Zeiss in 1911, this is a


modification by Paul Rudolph (1858-1935) and Charles Hastings
(1848-1932) of the symmetrical and monocentric triplet devised
by Steinheil in 1860. It is today one of the most common forms of
the loupe or pocket magnifier, but despite that humble occupation
it is close to diffraction limited and is the basic design of the TMB
Monocentric.

Erfle. Heinrich Erfle (1884-1923, pronounced Airfluh) was a


German physicist who worked for Steinheil & Shne and Carl
Zeiss. His design marks the transition to 20th century and the
first truly "wide field" eyepiece, which was not coincidentally
created for military applications (and was widely available as war
surplus in the 1950's). Patented in 1923, the three erfle
eyepieces consist of 5 elements in a 1-2-2 or 2-1-2 arrangement
that offers long eye relief but a relatively close alignment of field
lens to field stop. The distortion (for similar angular fields) is
comparable to that of the orthoscopic. This design has a 60
apparent field of view and eye relief of about 20mm (at
e = 28mm).
Kaspereit. The modification of the Erfle design by Otto Karl
Kaspereit adds a sixth lens to create a 2-2-2 design of three
corrected doublets. It is currently offered by Edmund Optics as the
"RKE Wide Field eyepiece", a design modified by David Rank. It
expands the apparent "wide field" of view to 68 and provides eye
relief of about 10mm (at e = 28mm).

Knig. Albert Knig (pronounced Kurnigk) patented in 1940.

Brandon. Chester Brandon designed military optics during World


War II and subsequently established his own optical
manufacturing company. The Brandon eyepiece resembles a
modified Plssl, but is made as two biconvex lenses of dense
barium crown capped by medium and high dispersion flints.
Brandons are widely esteemed by binary, planetary and lunar
astronomers for their central sharpness, high contrast, color
correction and lack of astigmatism and ghosting. As currently
manufactured by VernonScope, the lenses are fully coated but not
multicoated. The design has a 45 apparent field of view (not the
advertised 50) and eye relief of about 11mm (at e = 28mm).

Zeiss Astroplanokular. No longer manufactured, the Zeiss APO


is considered to be. This design has a 45 to 30 apparent field
of view and eye relief of about 11mm (at e = 28mm).

RKE. Trademarked in 1979 as the acronym for Rank Kaspereit


Erfle, "incorporating optical designs attributable to individuals
bearing those three names," the RKE eyepiece was designed for
Edmund Scientific Company by physicist David Rank (1907-1981).
It improves on the Kellner design by offering a wider, sharper
field, excellent aberration correction and longer eye relief. The
current design has a 45 apparent field of view and eye relief of
about 26mm (at e = 28mm). It is currently proprietary to
Edmund Optics.

Khler. This design was developed by Horst Khler for Zeiss in


1955. It is derived from the Erfle 2-1-2 design by splitting the
central positive lens into two asymmetrical lenses turned with
their stronger surfaces touching. Variants of this design were also
produced by Wright Scidmore and Hans Bertele, and later
manufactured by both TeleVue and Takahashi under the name
Panoptic. has a 25 to 30 apparent field of view and eye relief of
about 11mm (at e = 28mm).

Takahashi LE. The LE ("long eye relief") designs are


proprietary to Takahashi, This design has a 50 apparent field
of view and eye relief of about 11mm (at e = 28mm.

Nagler. Al Nagler was an optical designer for Farrand


and Keystone Camera who founded Tele Vue in 1977. In
the 1980's he patented and popularized a variety of
wide field eyepiece designs especially suitable for fast
focal ratio, hand guided Dobsonian telescopes
developed by John Dobson in the 1960s. Like many
wide field designs of the era, the Naglers are known for
their egregious weight and high cost (even in short focal
lengths), as well as their excellent correction of angular
magnification distortion, coma and spherical aberration.
The early Naglers suffered from severe spherical
aberration of the exit pupil (the so called "kidney bean"
blackout), which was largely but not entirely remedied
in later designs. This eyepiece has an 82 apparent field of view and eye relief of about 5mm (at
e = 13mm).

Pentax XW. The design here represents the 10 and 14mm


focal lengths. Compared to several other wide field designs
patented in the past few decades by Nagler, Koizumi, Kanai
and others, the innovation is the negative meniscus placed
immediately after the effective focal plane; this element is
moved ahead of the smyth doublet in the shorter focal lengths,
and replaces or is replaced by the smyth lens in the longest
focal lengths. All eyepieces in the series have complete
correction for coma, chromatic and spherical aberration, a 70
apparent field of view and a constant eye relief of 20mm. The
longest focal lengths have a slightly positive field curvature,
while the shortest focal lengths have a negative field
curvature.

Ethos. This design .

Pentax XW. The design here represents the 10 and 14mm


focal lengths. Compared to several other wide field designs
patented in the past few decades by Nagler, Koizumi, Kanai
and others, the innovation is the negative meniscus placed
immediately after the effective focal plane; this element is
moved ahead of the smyth doublet in the shorter focal
lengths, and replaces or is replaced by the smyth lens in the
longest focal lengths. All eyepieces in the series have
complete correction for coma, chromatic and spherical
aberration, a 70 apparent field of view and a constant eye
relief of 20mm. The longest focal lengths have a slightly
positive field curvature, while the shortest focal lengths have
a negative field curvature.
Ethos. This design .

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 2: Telescope & Eyepiece Combined - the optics of astronomical
telescopes.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic optical principles.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces, and results from my
collection.
"Eyepieces" - Chapter 37 in Herbert Gross, Fritz Blechinger & Bertram Achtner (eds.), Handbook of
Optical Systems, Volume 4: Survey of Optical Instruments. (Berlin: Wiley-VCH, 2008).
Optipedia - Online reference on optical topics by the SPIE, an international society advancing an
interdisciplinary approach to the science and application of light.

Optical Design for Visual Systems by Bruce H. Walker.


Lenses and Waves: Christiaan Huygens and the Mathematical Science of Optics - fine history of
early optical theory by Fokko Jan Dijksterhuis.
The Inverting Eyepiece and Its Evolution - a brief but useful summary of eyepiece design up to the
mid 20th century, by E. Wilfred Taylor.
The Evolution of the Eyepiece - Chris Lord's detailed narrative of the steps in astronomical eyepiece
development, not without factual errors (for example, "Max von Seidel" for Philipp Ludwig von Seidel).
The Amateur Astronomer's Handbook by J.B. Sidgwick.
Peter Dollond Answers Jesse Ramsden - An account of conflicting priority claims to the achromatic
doublet.
Eyepieces - John Savard's overview of eyepiece design with special attention to the selection of glasses.
Last revised 1/15/12 2012 Bruce MacEvoy

Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 2: Telescope & Eyepiece Combined - the optics of astronomical
telescopes.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic optical principles.
Astronomical Optics, Part 5: Evaluating Eyepieces - how to test eyepieces, and results from my
collection.
"Eyepieces" - Chapter 37 in Herbert Gross, Fritz Blechinger & Bertram Achtner (eds.), Handbook of
Optical Systems, Volume 4: Survey of Optical Instruments. (Berlin: Wiley-VCH, 2008).
Optipedia - Online reference on optical topics by the SPIE, an international society advancing an
interdisciplinary approach to the science and application of light.
Optical Design for Visual Systems by Bruce H. Walker.
Lenses and Waves: Christiaan Huygens and the Mathematical Science of Optics - fine history of
early optical theory by Fokko Jan Dijksterhuis.
The Inverting Eyepiece and Its Evolution - a brief but useful summary of eyepiece design up to the
mid 20th century, by E. Wilfred Taylor.
The Evolution of the Eyepiece - Chris Lord's detailed narrative of the steps in astronomical eyepiece
development, not without factual errors (for example, "Max von Seidel" for Philipp Ludwig von Seidel).
The Amateur Astronomer's Handbook by J.B. Sidgwick.
Peter Dollond Answers Jesse Ramsden - An account of conflicting priority claims to the achromatic
doublet.
Eyepieces - John Savard's overview of eyepiece design with special attention to the selection of glasses.
Last revised 1/15/12 2012 Bruce MacEvoy

type

standard

wide angle

super wide angle

apparent
diameter

diameter to
distance ratio

viewing distance

30

1:1.87

22.4 cm

40

1:1.37

16.4 cm

50

1:1.07

12.8 cm

60

1:0.87

10.4 cm

70

1:0.71

8.5 cm

80

1:0.60

7.2 cm

90

1:0.50

6.0 cm

100

1:0.42

5.0 cm

Astronomical Optics
Part 5: Evaluating Eyepieces
Once I had acquired a basic understanding of telescope optics and eyepiece design, I was ready to
make a much closer inspection of the eyepieces in my collection and intelligently evaluate the eyepieces
I borrowed from others.
There are many ways to evaluate eyepieces, and many criteria, but I was limited to the evaluations I
could make without an optical test or specific testing tools. Some of the tests could be objective (e.g.,
dimension measurement), but others (contrast, color) would be more or less subjective.
Eyepiece Virtues
If the mere reproduction of the image is accepted as a baseline performance standard, then eyepiece
virtues are the ways in which an ocular can qualitatively excel in that task. The four key virtues are
sharpness, brightness, contrast; certain comfort or convenience factors are important as well.
Sharpness is simply another word for veridical image reproduction: all elements of the image appear in
perfect focus, without any visible distortion or aberration. As all ocular errors increase with field height
(distance from the central optical axis), sharpness is typically evaluated as any degradation of the
image of a star as it is moved from the field center to the edge.
Brightness is the overall throughput or transmission efficiency of the eyepiece. Ideally, 100% of the
luminance of the image will be transmitted to the eye; the eyepiece will absorb or deflect no light. In
practice, brightness is reduced by the number of elements and groups (air/glass or glass/glass
boundaries) in an eyepiece, and by lenses that are uncoated. These all cause reflections and ghost
images, which divert light.
Contrast is the luminance difference between the lightest and darkest areas of an image, and is
primarily produced by high transmission oculars that do not scatter light. Light scattering is caused by
destructive interference in lens coatings, lens edges or ocular mountings that are not blackened.
The comfort and convenience factors in a good eyepiece will vary with the individual, but typically
include (1) comfortable eye relief, (2) eye guards or eye rests, especially in longer focal length
eyepieces, (3) size and weight, (4) apparent field of view.
Eyepiece Flaws
Eyepiece flaws occur when the ocular fails the minimum task of accurately reproducing the objective
image.
Scatter. Ideally, and separate from the errors of focus and projection, all light entering the eyepiece
should be transmitted to the focal plane as a coherent image. When this does not happen, a variety of
"stray light" artifacts appear in the image. These are reflections off interior surfaces that are combatted
by painting surfaces flat black, installing baffles, etc.
Ghosting is the appearance of secondary images of either the field stop or bright objects in the image.
The maximum number of potential ghost images is equal to the pairwise combination formula [N*(N
1)]/2, where N is the number of air/glass and glass/glass boundaries in the optical path that differ in
refractive index by more than 0.25. These are normally controlled by antireflection coatings, which are
very thin layers of materials with refractive indices that differ from each other and from air or glass by
less than 0.25, and therefore provide intermediate steps in the refraction.
Spherical Aberration of the Exit Pupil. If the exit rays of light leaving the eyepiece do not converge
in a single focal plane, there is no single, well defined location for the position of the observer's pupil.
This typically occurs because the peripheral rays are focused at a point closer to the eyepiece than the
central rays, which is a form of spherical aberration: thus, this defect is called spherical aberration of
the exit pupil, or SAEP. When this occurs, it produces an off axis "blackout" or blank area in the
eyepiece field of view opposite the direction of view, which has a lenticular form that is referred to as a
"kidney bean". Generally, observers will not object to an SAEP of 10% or less of the exit pupil diameter.

A. Manufacturing Quality
x. Packaging
x. End caps/cap fit
x. Field stop focus - sharply defined?
x. Internal blackening
x. Field fringe color - overcorrected = green or greenish blue, undercorrected = reddish, will not work
well with a fast objective.
x. Field illumination - evenly bright from center to edge: if not, and/or different locations are required to
see the field stop and to produce even illumination, then there is spherical aberration of the exit pupil.
x. Distortion - is the entire field stop visible without moving the eye? if not then there is either angular
or rectilinear distortion.
x. Scatter - Is the dark boundary of the field stop black, or milky? milky indicates scatter.
x. Mechanical and assembly quality
x. Binoviewer fit
x. Draw tube fit
x. Under cut
x. Weight
x. Scratch/dig ratio - Although optics can be laser tested, a measure of scratch width (in millionths or
106 millimeters) and pit diameter (in hundredths or 102 millimeters); often as high as 60/40 in
consumer optics but as low as 10/5 in military and certain industrial optics. Substitute glare, scatter
evaluations. Edmund Optics sells inexpensive, approximate and expensive, precise tools for assessing
scratch/dig in finished optics.
x. Coatings - Roland Christen evaluates eyepiece coatings by placing them in lateral daytime shade and
viewing them under sky light illumination with the black end lens caps on the field end. The photo
(below) shows the results using this method.

He also suggests checking the color of the coatings and glasses by viewing a white surface through the
eyepieces, as shown in the next photo (below).

B. Observing Comfort
x. Eye relief - Eye relief is usually shorter in higher power eyepieces. Some astronomers remedy short
eye relief in high power oculars by using lower power oculars with a barlow lens.
x. Exit pupil spherical aberration (dark adapted)
x. Critical alignment
x. Eye guard
x. Condensation
x. Parfocal range
C. Image Quality
x. Apparent field of View
The apparent field of view of an eyepiece is normally specified by the manufacturer, along with the
eyepiece focal length and design type. However, the nominal AFOV of an eyepiece is often inflated due
to the effects of positive distortion, especially in "wide field" eyepieces (AFOV > 60).
When the manufacturer's apparent field specification appears inaccurate, or is not known, then the
AFOV can be measured in three ways:
(1) Projective Measurement. Support the eyepiece between two stacks of books so that it is lying on its
side with the eye lens about 2 feet from a smooth wall or flat surface. Place a high powered flashlight
against the barrel so that the eyepiece projects the light from the flashlight onto the wall as a smooth
circle. (Try if possible to center the shadow of the flashlight bulb in the center of this circle.) Measure
the diameter of the circle, and the distance from the eyepiece mount to the wall. Then:
AFOV = 2*arctan(0.5*D/(PEP)).
(2) Physical Measurement. Use calipers to measure the interior diameter of the field stop, assuming it
exists and can be recognized. If in doubt, look through the eyepiece and use the point of a toothpick to
locate this as either the end of the eyepiece barrel or the interior edge of the lock nut holding the field
lens. (Note that many "super wide field" eyepieces locate the field stop inside the eyepiece where it
cannot be measured; in that case use 27 mm for eyepieces with a 1.25" barrel and 48 mm for
eyepieces with a 2" barrel.) Then the apparent field is simply the visual angle that the field stop

radius would subtend at 250 mm, divided by the apparent field radius . The distance size equation
then makes this equal to:
[15] AFOV = 2*arctan(e*(250/0.5*Dfs)).
For example, if the field stop is 23mm and the eyepiece focal length is 20mm, then:
AFOV = 2*arctan(e*(250/0.5*Dfs)) = 2*arctan(20*(250/0.5*23) = 2*arctan(434.8) = 2*89.9 = .
(3) Visual Measurement. Tape to a convenient wall a long strip of paper or tape measure on which small
increments (inches or centimeters) are clearly marked and visible from a distance (Dm) of about 100
cm. Hold the eyepiece over one eye so that the full field is clearly visible and superimposed over the
measurement tape, visible with the other eye. Manipulate the eyepiece position until the tape appears
to measure the full diameter of the eyepiece field (De). Then:
[16] AFOV = 2*arctan(0.5*De/Dm).
For example, viewing a metric measuring tape from 100 cm, you measure an apparent width of the
eyepiece field as 93 cm. Then:
[17] AFOV = 2*arctan(0.5*93/100) = 2*arctan(0.47) = 2*25 = 50
x. True Field of View
(Multiplication by 60 converts degrees to arcminutes.) If the field stop diameter and objective focal
length are known, then it can be calculated as:
[18] TFOV = Dfs/o*57.3.
This formula also shows that the maximum true field of view possible with a 1.25" eyepiece barrel or a
visual back of a certain diameter. For a 10" /8 telescope, the maximum true field of view is:
TFOV = 27/2032*57.3. = 0.76
If the field stop diameter and apparent field of view are not known, the true field of view can be
measured by means of star transit times. This requires at least three measurements of the time (in
seconds) it takes a bright star, placed outside the eyepiece field of view, to appear at one side of the
eyepiece field, cross the center of the field, and disappear at the opposite side. Care must be taken to
hold the head in a fixed position, and to pass the star through the center of the eyepiece field. Then the
true field is the average of the three timings multiplied by the cosine of the star's declination (stars
closer to the celestial pole will require a longer time to transit the eyepiece field):
[19] TFOV = 0.25*cos(Declstar)*t.
(The factor of 0.25 is necessary to convert from seconds in time to arcminutes of angular width.) For
example, using Regulus (alpha Leonis) as the transit star (Declination = +1158'), three eyepiece
transit times are 134, 131 and 138 seconds. Then:
TFOV = 0.25*cos(11.97)*(134+131+138)/3 = 0.25 * 0.978 * 134.3 = 32.8 arcminutes.
Given the TFOV (in arcminutes) and system magnification, the AFOV (in degrees) is:
[20] AFOV = TFOV/Mt * 60.
The TFOV determined by star transit measurements will typically not correspond exactly to the TFOV
calculated from the manufacturer supplied eyepiece apparent field of view, and the magnification
calculated from the manufacturer supplied eyepiece focal length. This is usually because the eyepiece
AFOV and e are nominal, inferred from the computer optical design rather than the physical
construction of the eyepiece; and because timing star transits includes some measurement error.
x. True field accuracy/variation within the line
x. Color - Two excellent tests are (a) moon, (b) WZ Cas
x. Contrast - (a) moon, (b) jupiter, (c) deep sky (illumination level, and pupil size) - faintest visible
extended area on the moon.
x. Transmission - faintest visible stars
x. Diffusion - (a) center field, (b) bar, (c) field edge
x. Glare, Scatter - (a) off field sirius, (b) center field sirius
x. Ghosting - how bright? colored? follow, mirror image, or remain in center? in focus or diffuse?
x. Barlow image
x. Spherical aberration
x. Astigmatism - worse with wider field and faster objective
x. Coma - worse with wider field and faster objective
x. Field curvature - (center to edge focus, amount of focuser turn) ...
x. Angular magnification distortion - requires rectilinear pattern
x. Rectilinear distortion - requires rectilinear pattern
x. Lateral color - in almost all eyepieces. increase near edge of field? red toward center
undercorrected, blue toward center overcorrected.
x. Lateral color

The Eyepieces
These were tested using the above criteria.
BAADER PLANETARIUM
These are the most expensive wide field eyepieces that I own.
DENKMEIER
These are the most expensive wide field eyepieces that I own.
EDMUND OPTICS
I purchased this simple magnifier or "loupe" lens because the early monocentric triplets were highly favored by late 19th century
astronomers.
Hastings Triplet
This is that.
RKE (Reverse Kellner Eyepiece)
This is that.
EXPLORE SCIENTIFIC
These eyepieces also have a very strong reputation.
MEADE
The RKE design, like the Brandon, has a very high reputation among planetary and lunar observers, historical problems with
quality control notwithstanding.
PENTAX
On disassembly, the Meade 5000 Plssl eyepieces turn out to be the Erfle II 2-1-2 design.
RUSSELL OPTICS
These are the most expensive wide field eyepieces that I own.
TAKAHASHI
Edmund Optics sells this as a single eyepiece.
The University Optics planetary eyepiece is a modified erfle design.
TELE VUE
The Plssl is a straightforward 2-2 design
The Tele Vue plssls
TMB OPTICAL
These are the most expensive wide field eyepieces that I own.
UNIVERSITY OPTICS
The original Abbe ortho design is considered one of the most important eyepiece designs of the last two centuries and is
remarkable for its nearly flat field, free of distortion.
VERNONSCOPE
The Brandon design eyepieces, now manufactured by VernonScope, have one of the strongest reputations among planetary
observers for their brightness, contrast and flat field.
VIXEN
Further Reading
Astronomical Optics, Part 1: Basic Optics - an overview of basic optics.
Astronomical Optics, Part 2: Telescope & Eyepiece Combined - the optics of astronomical telescopes.
Astronomical Optics, Part 3: Optical Aberrations - an overview of basic optical principles.
Astronomical Optics, Part 4: Eyepiece Designs - an illustrated overview of historically important eyepiece designs.
2011 Bruce MacEvoy

Das könnte Ihnen auch gefallen