02.02.2015 Views

Gathering light; the telescope - University of Arizona

Gathering light; the telescope - University of Arizona

Gathering light; the telescope - University of Arizona

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

2.1.Basic Principles<br />

Chapter 2: <strong>Ga<strong>the</strong>ring</strong> <strong>light</strong> - <strong>the</strong> <strong>telescope</strong><br />

Astronomy centers on <strong>the</strong> study <strong>of</strong> vanishingly faint signals, <strong>of</strong>ten from complex fields <strong>of</strong><br />

sources. Job number one is <strong>the</strong>refore to collect as much <strong>light</strong> as possible, with <strong>the</strong> highest<br />

possible angular resolution. So life is simple. We want to: 1.) build <strong>the</strong> largest <strong>telescope</strong><br />

we can afford (or can get someone else to buy for us), 2.) design it to be efficient and 3.)<br />

at <strong>the</strong> same time shield <strong>the</strong> signal from unwanted contamination, 4.) provide diffractionlimited<br />

images over as large an area in <strong>the</strong> image plane as we can cover with detectors,<br />

and 5.) adjust <strong>the</strong> final beam to match <strong>the</strong> signal optimally onto those detectors. Figure<br />

2.1 shows a basic <strong>telescope</strong> that we might use to achieve <strong>the</strong>se goals.<br />

Achieving<br />

condition 4.) is a<br />

very strong driver<br />

on <strong>telescope</strong><br />

design. The<br />

etendue is constant<br />

only for a beam<br />

passing through a<br />

perfect optical<br />

system. For such a<br />

system, <strong>the</strong> image<br />

Figure 2.1. Simple Prime-Focus Telescope<br />

FoTelescope primary mirror parameters.<br />

quality is limited by <strong>the</strong> wavelength <strong>of</strong> <strong>the</strong> <strong>light</strong> and <strong>the</strong> diffraction at <strong>the</strong> limiting<br />

aperture for <strong>the</strong> <strong>telescope</strong> (normally <strong>the</strong> edge <strong>of</strong> <strong>the</strong> primary mirror). Assuming that <strong>the</strong><br />

<strong>telescope</strong> primary is circular, <strong>the</strong> resulting image illumination is<br />

where m = (π r 0 θ / λ), r 0 is <strong>the</strong> radius<br />

<strong>of</strong> <strong>the</strong> <strong>telescope</strong> aperture (=D/2 in<br />

Figure 2.1), is <strong>the</strong> wavelength <strong>of</strong><br />

operation, and J 1 is a Bessel function<br />

<strong>of</strong> <strong>the</strong> first kind. The result is <strong>the</strong> wellknown<br />

Airy function (Figure 2.2),<br />

named after <strong>the</strong> British astronomer<br />

George Biddell Airy. We will illustrate<br />

a simple derivation later in this<br />

chapter.<br />

The first zero <strong>of</strong> <strong>the</strong> Bessel function<br />

occurs at m= 1.916, or for small θ, at θ Figure 2.2. The Airy Function<br />

1.22 λ / D. The classic Rayleigh<br />

Criterion states that two point sources <strong>of</strong> equal brightness can be distinguished only if<br />

<strong>the</strong>ir separation is > 1.22 /D; that is, <strong>the</strong> peak <strong>of</strong> one image is no closer than <strong>the</strong> first<br />

dark ring in <strong>the</strong> o<strong>the</strong>r. With modern image analysis techniques, this criterion can be


eadily surpassed. The full width at half maximum <strong>of</strong> <strong>the</strong> central image is about 1.03 /D<br />

for an unobscured aperture and becomes s<strong>light</strong>ly narrower if <strong>the</strong>re is a blockage in <strong>the</strong><br />

center <strong>of</strong> <strong>the</strong> aperture (e.g., a secondary mirror).<br />

Although we will emphasize a more ma<strong>the</strong>matical approach in this chapter, recall from<br />

<strong>the</strong> previous chapter that <strong>the</strong> diffraction pattern can be understood as a manifestation <strong>of</strong><br />

Fermat’s Principle. Where <strong>the</strong> ray path lengths are so nearly equal that <strong>the</strong>y interfere<br />

constructively, we get <strong>the</strong> central image. The first dark ring lies where <strong>the</strong> rays are out <strong>of</strong><br />

phase and interfere destructively, and <strong>the</strong><br />

remainder <strong>of</strong> <strong>the</strong> <strong>light</strong> and dark rings represent<br />

constructive and destructive interference at<br />

increasing path differences.<br />

Although <strong>the</strong> diffraction limit is <strong>the</strong> ideal, practical<br />

optics have a series <strong>of</strong> shortcomings, described as<br />

aberrations. There are three primary geometric<br />

aberrations:<br />

1. spherical, occurs when an <strong>of</strong>f-axis input ray is<br />

directed in front <strong>of</strong> or behind <strong>the</strong> image<br />

position for an on-axis input<br />

ray, with rays at <strong>the</strong> same <strong>of</strong>faxis<br />

angle crossing <strong>the</strong> image<br />

plane symmetrically<br />

distributed around <strong>the</strong> on-axis<br />

image. Spherical aberration<br />

tends to yield a blurred halo<br />

around an image (Figure 2.3).<br />

Figure 2.3. The most famous example<br />

<strong>of</strong> spherical aberration, <strong>the</strong> Hubble<br />

Space Telescope before various means<br />

were taken to correct <strong>the</strong> problem.<br />

A sphere forms a perfect Figure 2.4. The focusing properties <strong>of</strong> a spherical reflector.<br />

image <strong>of</strong> a point source<br />

located at <strong>the</strong> center <strong>of</strong> <strong>the</strong> sphere, with <strong>the</strong> image produced on top <strong>of</strong> <strong>the</strong> source.<br />

However, for a source at infinity, <strong>the</strong> distance, F, where <strong>the</strong> reflected ray crosses <strong>the</strong><br />

axis <strong>of</strong> <strong>the</strong> mirror is<br />

where R is <strong>the</strong> radius <strong>of</strong> curvature <strong>of</strong> <strong>the</strong> mirror and is <strong>the</strong><br />

angle <strong>of</strong> reflectance from <strong>the</strong> surface <strong>of</strong> <strong>the</strong> mirror. That is, <strong>the</strong><br />

reflected rays cross <strong>the</strong> mirror axis at smaller values <strong>of</strong> F <strong>the</strong><br />

far<strong>the</strong>r <strong>of</strong>f-axis <strong>the</strong>y impinge on <strong>the</strong> mirror. The result is<br />

shown in Figure 2.4; <strong>the</strong>re is no point along <strong>the</strong> optical axis<br />

with a well-formed image.<br />

2. coma occurs when input rays arriving at an angle from <strong>the</strong><br />

optical axis miss toward <strong>the</strong> same side <strong>of</strong> <strong>the</strong> on-axis image no<br />

matter where <strong>the</strong>y enter <strong>the</strong> <strong>telescope</strong> aperture, and with a<br />

Figure 2.5. Comatic<br />

image displayed to<br />

show interference.


progressive increase in image diameter with increasing distance from <strong>the</strong> center <strong>of</strong> <strong>the</strong><br />

field. Coma is axially symmetric in <strong>the</strong> sense that similar patterns are generated by<br />

input rays at <strong>the</strong> same <strong>of</strong>f axis angle at all azimuthal positions. Comatic images have<br />

characteristic fan or comet-shapes with<br />

<strong>the</strong> tail pointing away from <strong>the</strong> center <strong>of</strong><br />

<strong>the</strong> image plane (Figure 2.5).<br />

A paraboloidal reflector provides an<br />

example. The paraboloid by construction<br />

produces a perfect image <strong>of</strong> a point<br />

source on <strong>the</strong> axis <strong>of</strong> <strong>the</strong> mirror and at<br />

infinite distance. As shown in Figure 2.6,<br />

<strong>the</strong> focus for a ray that is <strong>of</strong>f-axis by <strong>the</strong><br />

angle is displaced from <strong>the</strong> axis <strong>of</strong> <strong>the</strong><br />

mirror by<br />

<br />

3. astigmatism (Figure 2.7) is a cylindrical<br />

wavefront distortion resulting from an<br />

optical system that has different focal<br />

planes for an <strong>of</strong>f-axis object in one<br />

direction from <strong>the</strong> optical axis <strong>of</strong> <strong>the</strong><br />

system compared with <strong>the</strong> orthogonal direction. It results in images that are elliptical<br />

on ei<strong>the</strong>r side <strong>of</strong> best-focus, with <strong>the</strong><br />

direction <strong>of</strong> <strong>the</strong> long axis <strong>of</strong> <strong>the</strong> ellipse<br />

changing by 90 degrees going from<br />

ahead to behind focus.<br />

Two o<strong>the</strong>r aberrations are less fundamental<br />

but in practical systems can degrade <strong>the</strong><br />

entendue:<br />

4. curvature <strong>of</strong> field, which occurs when<br />

<strong>the</strong> best images are not formed at a<br />

plane but instead on a surface that is<br />

convex or concave toward <strong>the</strong> <strong>telescope</strong><br />

entrance aperture (see Figure 2.8).<br />

Figure 2.6. Focusing properties <strong>of</strong> a<br />

paraboloidal mirror.<br />

5. distortion arises when <strong>the</strong> image scale<br />

changes over <strong>the</strong> focal plane; that is, if a<br />

set <strong>of</strong> point sources placed on a uniform<br />

grid is observed, <strong>the</strong>ir relative image<br />

positions are displaced from <strong>the</strong><br />

corresponding grid positions at <strong>the</strong> focal<br />

plane. Figure 2.9 illustrates symmetric<br />

Figure 2.7. Illustration <strong>of</strong> astigmatism (from<br />

Starizona). If we assume <strong>the</strong> astigmatism is<br />

due to an asymmetry in <strong>the</strong> optics, <strong>the</strong>n <strong>the</strong><br />

tangential plane contains both <strong>the</strong> object being<br />

imaged and <strong>the</strong> axis <strong>of</strong> symmetry, and <strong>the</strong><br />

sagittal plane is orthogonal to <strong>the</strong> tangential<br />

one.


forms <strong>of</strong> distortion, but this defect<br />

can occur in a variety <strong>of</strong> o<strong>the</strong>r forms,<br />

such as trapezoidal.<br />

Nei<strong>the</strong>r <strong>of</strong> <strong>the</strong>se latter two aberrations<br />

actually degrade <strong>the</strong> images <strong>the</strong>mselves.<br />

The effects <strong>of</strong> field curvature can be<br />

removed by suitable design <strong>of</strong> <strong>the</strong><br />

following optics, or by curving <strong>the</strong><br />

detector surface to match <strong>the</strong> curvature Figure 2.8. Curvature <strong>of</strong> field<br />

<strong>of</strong> field. Distortion can be removed by<br />

resampling <strong>the</strong> images and placing <strong>the</strong>m at <strong>the</strong> correct positions, based on a previous<br />

characterization <strong>of</strong> <strong>the</strong> distortion effects. Such resampling may degrade <strong>the</strong> images,<br />

particularly if <strong>the</strong>y are poorly sampled (e.g., <strong>the</strong> detectors are too large to return all <strong>the</strong><br />

information about <strong>the</strong> images) but need not if care is used both in <strong>the</strong> system design and<br />

<strong>the</strong> method to remove <strong>the</strong> distortion.<br />

Optical systems using lenses are also<br />

subject to:<br />

6. chromatic aberration, resulting<br />

when <strong>light</strong> <strong>of</strong> different colors is<br />

not brought to <strong>the</strong> same focus.<br />

A central aspect <strong>of</strong> optics design is Figure 2.9. Distortion.<br />

that aberrations introduced by an<br />

optical element can be compensated with a following element to improve <strong>the</strong> image<br />

quality in multi-element optical systems. For example, some high performance<br />

commercial camera lenses have up to 20 elements that all work toge<strong>the</strong>r to provide high<br />

quality images over a large field.<br />

In addition to <strong>the</strong>se aberrations, <strong>the</strong> <strong>telescope</strong> performance can be degraded by<br />

manufacturing errors that result in <strong>the</strong> optical elements not having <strong>the</strong> exact prescribed<br />

shapes, by distortions <strong>of</strong> <strong>the</strong> elements in <strong>the</strong>ir mountings, and by mis-alignments.<br />

Although <strong>the</strong>se latter effects are also sometimes termed aberrations, <strong>the</strong>y are <strong>of</strong> a<br />

different class from <strong>the</strong> six aberrations we have listed and we will avoid this terminology.<br />

In operation, <strong>the</strong> images can be fur<strong>the</strong>r degraded by atmospheric seeing, <strong>the</strong> disturbance<br />

<strong>of</strong> <strong>the</strong> wavefronts <strong>of</strong> <strong>the</strong> <strong>light</strong> from <strong>the</strong> source as <strong>the</strong>y pass through a turbulent atmosphere<br />

with refractive index variations due to temperature inhomogeneity. A rough<br />

approximation <strong>of</strong> <strong>the</strong> behavior is that <strong>the</strong>re are atmospheric bubbles <strong>of</strong> size r 0 = 5 – 15 cm<br />

with temperature variations <strong>of</strong> a few hundredths up to 1 o C, moving at wind velocities <strong>of</strong><br />

10 to 50 m/sec (r 0 is defined by <strong>the</strong> typical size effective at a wavelength <strong>of</strong> 0.5m and<br />

called <strong>the</strong> Fried parameter). The time scale for variations over a typical size <strong>of</strong> r 0 at <strong>the</strong><br />

<strong>telescope</strong> is <strong>the</strong>refore <strong>of</strong> order 10msec. For a <strong>telescope</strong> with aperture smaller than r 0 , <strong>the</strong><br />

effect is to cause <strong>the</strong> images formed by <strong>the</strong> <strong>telescope</strong> to move as <strong>the</strong> wavefronts are tilted


to various angles by <strong>the</strong> passage <strong>of</strong> warmer and cooler air bubbles. If <strong>the</strong> <strong>telescope</strong><br />

aperture is much larger than r 0 , many different r 0 -sized columns are sampled at once.<br />

Images taken over significantly longer than 10msec are called seeing-limited, and have<br />

typical sizes <strong>of</strong> /r 0 , derived similarly to equation (1.12) since <strong>the</strong> wavefront is preserved<br />

accurately only over a patch <strong>of</strong> diameter ~ r 0 . These images may be 0.5 to 1 arcsec in<br />

diameter, or larger under poor conditions. However, since <strong>the</strong> phase <strong>of</strong> <strong>the</strong> <strong>light</strong> varies<br />

quickly over each r 0 -diameter patch, a complex and variable interference pattern is<br />

formed at <strong>the</strong> focal plane due to <strong>the</strong> interference among <strong>the</strong>se different patches. A fast<br />

exposure (e.g., ~ 10msec) freezes this pattern and <strong>the</strong> image appears speckled, within <strong>the</strong><br />

overall envelope <strong>of</strong> <strong>the</strong> seeing limit. This behavior is strongly wavelength-dependent.<br />

Because <strong>the</strong> refractive index <strong>of</strong> air decreases with increasing wavelength, r 0 increases<br />

roughly as 6/5 , so, for example, <strong>the</strong> effect <strong>of</strong> seeing at 10m is largely image motion<br />

even with a 10-m <strong>telescope</strong>. These topics will be covered more extensively in Chapter 7.<br />

There are a number <strong>of</strong> ways to describe <strong>the</strong> imaging performance <strong>of</strong> a <strong>telescope</strong> or o<strong>the</strong>r<br />

optical system. The nature <strong>of</strong> <strong>the</strong> diffraction-limited image is a function <strong>of</strong> <strong>the</strong> <strong>telescope</strong><br />

configuration (shape <strong>of</strong> primary mirror, size <strong>of</strong> secondary mirror, etc.), so we define<br />

“perfect” imaging as that obtained for <strong>the</strong> given configuration and with all o<strong>the</strong>r aspects<br />

<strong>of</strong> <strong>the</strong> <strong>telescope</strong> performing perfectly. The ratio <strong>of</strong> <strong>the</strong> signal from <strong>the</strong> brightest point in<br />

this perfect image divided into <strong>the</strong> signal from <strong>the</strong> same point in <strong>the</strong> achieved image is <strong>the</strong><br />

Strehl ratio. We can also measure <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> total energy in an image to <strong>the</strong> energy<br />

received through a given round aperture at <strong>the</strong> focal plane, a parameter called <strong>the</strong><br />

encircled energy. We can also describe <strong>the</strong> root-mean-square (rms) errors <strong>of</strong> <strong>the</strong> delivered<br />

wavefronts compared with those for <strong>the</strong> “perfect” image. There is no simple way to relate<br />

<strong>the</strong>se performance descriptors to one ano<strong>the</strong>r, since <strong>the</strong>y each depend on different aspects<br />

<strong>of</strong> <strong>the</strong> optical performance. However, roughly speaking, a <strong>telescope</strong> can be considered to<br />

be diffraction limited if <strong>the</strong> Strehl ratio is > 0.8 or <strong>the</strong> rms wavefront errors are < /14,<br />

where is <strong>the</strong> wavelength <strong>of</strong> observation. This relation is known as <strong>the</strong> Maréchal<br />

criterion. A useful approximate relation between <strong>the</strong> Strehl ratio, S, and <strong>the</strong> rms<br />

wavefront error, , is also an extended version <strong>of</strong> an approximation due to Maréchal:<br />

<br />

2.2.Telescope design<br />

We will discuss general <strong>telescope</strong> design in <strong>the</strong> context <strong>of</strong> optical <strong>telescope</strong>s in <strong>the</strong><br />

following two sections, and <strong>the</strong>n in Section 2.4 will expand <strong>the</strong> treatment to <strong>telescope</strong>s<br />

for o<strong>the</strong>r wavelength regimes. In general, <strong>telescope</strong>s require <strong>the</strong> use <strong>of</strong> combinations <strong>of</strong><br />

multiple optical elements. The basic formulae we will use for simple geometric optics can<br />

be found in Section 1.3 (in Chapter 1). For more, consult any optics text.<br />

However, all <strong>of</strong> <strong>the</strong>se relations are subsumed for modern <strong>telescope</strong> design into computer<br />

“ray-tracing” programs that follow multiple rays <strong>of</strong> <strong>light</strong> through a hypo<strong>the</strong>tical train <strong>of</strong><br />

optical elements, applying <strong>the</strong> laws <strong>of</strong> refraction and reflection at each surface, to<br />

produce a simulated image. The properties <strong>of</strong> this image can be optimized iteratively,<br />

with assistance from <strong>the</strong> program. These programs usually can also include diffraction to<br />

provide a physical optics output in addition to <strong>the</strong> geometric ray trace.


None<strong>the</strong>less, <strong>the</strong>re are basic parameters in a <strong>telescope</strong> and in o<strong>the</strong>r optical systems that<br />

you need to understand. We need to define some terms before discussing <strong>the</strong>m:<br />

1. A real image is where <strong>the</strong> <strong>light</strong> from an unresolved source is brought to its greatest<br />

concentration – putting a detector at <strong>the</strong> position <strong>of</strong> <strong>the</strong> real image will record <strong>the</strong><br />

source.<br />

2. A virtual image can be formed if <strong>the</strong> optics produce a bundle <strong>of</strong> <strong>light</strong> that appears to<br />

come from a real image but no such image has been formed – for example, a convex<br />

mirror may reflect <strong>light</strong> so it appears that <strong>the</strong>re is a real image behind <strong>the</strong> mirror.<br />

Except for very specialized applications, refractive <strong>telescope</strong>s using lenses for <strong>the</strong>ir<br />

primary <strong>light</strong> collectors are no longer employed in astronomy; in addition to chromatic<br />

aberration, large lenses are difficult to support without distortion and <strong>the</strong>y absorb some <strong>of</strong><br />

<strong>the</strong> <strong>light</strong> ra<strong>the</strong>r than delivering it to <strong>the</strong> image plane (in <strong>the</strong> infrared, virtually all <strong>the</strong> <strong>light</strong><br />

is absorbed). However, refractive elements are important in o<strong>the</strong>r aspects <strong>of</strong> advanced<br />

<strong>telescope</strong> design, to be discussed later.<br />

Reflecting <strong>telescope</strong>s are achromatic and are also efficient because <strong>the</strong>y can be provided<br />

with high efficiency mirror surfaces. From <strong>the</strong> radio to <strong>the</strong> near-ultraviolet, metal mirror<br />

surfaces provide reflection efficiencies > 90%, and usually even > 95%. At wavelengths<br />

shorter than ~ 0.15m, metals have poor reflection and multilayer stacks <strong>of</strong> dielectrics are<br />

used, but with efficiencies <strong>of</strong> only 50 – 70%. Significantly short <strong>of</strong> 0.01m (energy > 0.1<br />

keV), grazing incidence reflection can still be used for <strong>telescope</strong>s, but with significant<br />

changes in configuration as discussed in Section 2.4.4.<br />

Although modern computer-aided optical design would in principle allow a huge<br />

combination <strong>of</strong> two-mirror <strong>telescope</strong> systems, <strong>the</strong> traditional conic-section-based<br />

concepts already work very well. The ability <strong>of</strong> conic section mirrors to form images is<br />

easily described in terms <strong>of</strong> Fermat’s Principle. As <strong>the</strong> surface <strong>of</strong> a sphere is everywhere<br />

<strong>the</strong> same distance from its center, so a concave spherical mirror images an object onto<br />

itself when that object is placed at <strong>the</strong> center <strong>of</strong> curvature. An ellipse defines <strong>the</strong> surface<br />

that has a constant sum <strong>of</strong> <strong>the</strong> distances from one focus to ano<strong>the</strong>r, so an ellipsoid images<br />

an object at one focus at <strong>the</strong> position <strong>of</strong> its second focus. A parabola can be described as<br />

an ellipse with one focus moved to infinite distance, so a paraboloid images an object at<br />

large distant to its focus. A hyperbola is defined as <strong>the</strong> locus <strong>of</strong> points whose difference<br />

<strong>of</strong> distances from two foci is constant. Thus, a hyperboloid forms a perfect image <strong>of</strong> a<br />

virtual image.<br />

The basic <strong>telescope</strong> types based on <strong>the</strong> properties <strong>of</strong> <strong>the</strong>se conic sections are:<br />

1. A prime focus <strong>telescope</strong> (Figure 2.1) has a paraboloidal primary mirror and forms<br />

images directly at <strong>the</strong> mirror focus; since <strong>the</strong> images are in <strong>the</strong> center <strong>of</strong> <strong>the</strong><br />

incoming beam <strong>of</strong> <strong>light</strong>, this arrangement is inconvenient unless <strong>the</strong> <strong>telescope</strong> is<br />

so large that <strong>the</strong> detector receiving <strong>the</strong> <strong>light</strong> blocks a negligible portion <strong>of</strong> this<br />

beam


2. A Newtonian <strong>telescope</strong> (Figure 2.10) uses a flat mirror tilted at 45 o to bring <strong>the</strong><br />

focus to <strong>the</strong> side <strong>of</strong> <strong>the</strong> incoming beam <strong>of</strong> <strong>light</strong> where it is generally more<br />

conveniently accessed<br />

3. A Gregorian <strong>telescope</strong> brings <strong>the</strong> <strong>light</strong> from its paraboloidal primary mirror to a<br />

focus, and <strong>the</strong>n uses an ellipsoidal mirror beyond this focus to bring it to a second<br />

focus. Usually this <strong>light</strong> passes through a hole in <strong>the</strong> center <strong>of</strong> <strong>the</strong> primary and <strong>the</strong><br />

second focus is conveniently behind <strong>the</strong> primary and out <strong>of</strong> <strong>the</strong> incoming beam.<br />

4. A Cassegrain <strong>telescope</strong> intercepts <strong>the</strong> <strong>light</strong> from its paraboloidal primary ahead <strong>of</strong><br />

<strong>the</strong> focus with a convex hyperboloidal mirror. This mirror re-focuses <strong>the</strong> <strong>light</strong><br />

from <strong>the</strong> virtual image formed by <strong>the</strong> primary to a second focus, usually behind<br />

<strong>the</strong> primary mirror as in <strong>the</strong> Gregorian design.<br />

Assuming perfect<br />

manufacturing <strong>of</strong><br />

<strong>the</strong>ir optics and<br />

maintenance <strong>of</strong><br />

<strong>the</strong>ir alignment,<br />

all <strong>of</strong> <strong>the</strong>se<br />

<strong>telescope</strong> types<br />

are limited in<br />

image quality by<br />

<strong>the</strong> coma due to<br />

<strong>the</strong>ir paraboloidal<br />

primary mirrors.<br />

Some less obvious<br />

conic section<br />

combinations are<br />

<strong>the</strong> Dall-Kirkham,<br />

with an ellipsoidal<br />

primary and<br />

spherical<br />

secondary, and <strong>the</strong><br />

inverted Dall-<br />

Kirkham, with a<br />

spherical primary<br />

and ellipsoidal<br />

secondary. Both<br />

suffer from larger<br />

amounts <strong>of</strong> coma<br />

than for <strong>the</strong><br />

paraboloidprimary<br />

types, and<br />

hence <strong>the</strong>y are<br />

seldom used.<br />

Figure 2.10. The basic optical <strong>telescope</strong> types.


2.3.Matching Telescopes to Instruments<br />

2.3.1. Telescope Parameters<br />

To match <strong>the</strong> <strong>telescope</strong> output to an instrument, we need to describe <strong>the</strong> emergent beam<br />

<strong>of</strong> <strong>light</strong> in detail. During <strong>the</strong> process <strong>of</strong> designing an instrument, usually an optical model<br />

<strong>of</strong> <strong>the</strong> <strong>telescope</strong> is included in <strong>the</strong> ray trace <strong>of</strong> <strong>the</strong> instrument to be sure that <strong>the</strong> two work<br />

as expected toge<strong>the</strong>r. A convenient vocabulary to describe <strong>the</strong> general matching <strong>of</strong> <strong>the</strong><br />

two is based on <strong>the</strong> following terms:<br />

The focal length is measured by projecting <strong>the</strong> conical bundle <strong>of</strong> rays that arrives at <strong>the</strong><br />

focus back until its diameter matches <strong>the</strong> aperture <strong>of</strong> <strong>the</strong> <strong>telescope</strong>; <strong>the</strong> focal length is <strong>the</strong><br />

length <strong>of</strong> <strong>the</strong> resulting projected bundle <strong>of</strong> rays. See Figure 2.1 for <strong>the</strong> simplest example.<br />

This definition allows for <strong>the</strong> effects, for example, <strong>of</strong> secondary mirrors in modifying <strong>the</strong><br />

properties <strong>of</strong> <strong>the</strong> ray bundle produced by <strong>the</strong> primary mirror.<br />

The f-number <strong>of</strong> <strong>the</strong> <strong>telescope</strong> is <strong>the</strong> diameter <strong>of</strong> <strong>the</strong> incoming ray bundle, i.e. <strong>the</strong><br />

<strong>telescope</strong> aperture, divided into <strong>the</strong> focal length, f/D in Figure 2.1. This term is also used<br />

as a general description <strong>of</strong> <strong>the</strong> angle <strong>of</strong> convergence <strong>of</strong> a beam <strong>of</strong> <strong>light</strong>, in which case it is<br />

<strong>the</strong> diameter <strong>of</strong> <strong>the</strong> beam at some distance from its focus, divided into this distance. The<br />

f-number can also be expressed as 0.5 cotan (θ), where θ is <strong>the</strong> beam half-angle. Beams<br />

with large f-numbers are “slow” while those with small ones are “fast.”<br />

Equation (1.11) can be used to determine <strong>the</strong> f-number <strong>of</strong> <strong>the</strong> beam at <strong>the</strong> focus <strong>of</strong> a<br />

<strong>telescope</strong>, given <strong>the</strong> desired mapping <strong>of</strong> <strong>the</strong> size <strong>of</strong> a detector element into its resolution<br />

on <strong>the</strong> sky. From equation (1.11), we can match to any detector area by adjusting <strong>the</strong><br />

optics to shape <strong>the</strong> beam to an appropriate solid angle. Simplifying to a round <strong>telescope</strong><br />

aperture <strong>of</strong> diameter D and a detector <strong>of</strong> diameter d,<br />

D d<br />

in<br />

out<br />

Thus, if <strong>the</strong> detector element is 10m = 1 X 10 -5 m on a side and we want it to map to 0.1<br />

arcsec = 4.85 X 10 -7 rad on <strong>the</strong> sky on a 10-m aperture <strong>telescope</strong>, we have<br />

7<br />

5<br />

10m<br />

4.8510<br />

rad 110<br />

m<br />

out<br />

Solving for θ out , <strong>the</strong> f-number <strong>of</strong> <strong>the</strong> beam into <strong>the</strong> detector is 1.0, a very fast beam<br />

indeed. This calculation previews an issue to be discussed in more detail later, that <strong>of</strong><br />

matching <strong>the</strong> outputs <strong>of</strong> large <strong>telescope</strong>s onto modern detector arrays with <strong>the</strong>ir relatively<br />

small pixels.<br />

The plate scale is <strong>the</strong> translation from physical units, e.g. mm, to angular ones, e.g.<br />

arcsec, at <strong>the</strong> <strong>telescope</strong> focal plane. As for <strong>the</strong> derivation <strong>of</strong> <strong>the</strong> thin lens formula, we can<br />

determine <strong>the</strong> plate scale by making use <strong>of</strong> <strong>the</strong> fact that any ray passing through <strong>telescope</strong><br />

on <strong>the</strong> optical axis <strong>of</strong> its equivalent single lens is not deflected. Thus, if we define <strong>the</strong><br />

magnification <strong>of</strong> <strong>the</strong> <strong>telescope</strong>, m, to be <strong>the</strong> ratio <strong>of</strong> <strong>the</strong> f number delivered to <strong>the</strong> focal<br />

plane to <strong>the</strong> primary mirror f number, <strong>the</strong>n <strong>the</strong> equivalent focal length <strong>of</strong> <strong>the</strong> <strong>telescope</strong> is<br />

and <strong>the</strong> angle on <strong>the</strong> sky corresponding to a distance b at <strong>the</strong> focal plane is


The field <strong>of</strong> view (FOV) is <strong>the</strong> total angle on <strong>the</strong> sky that can be imaged by <strong>the</strong> <strong>telescope</strong>.<br />

It might be limited by <strong>the</strong> <strong>telescope</strong> optics and baffles, or by <strong>the</strong> dimensions <strong>of</strong> <strong>the</strong><br />

detector (or acceptance angle <strong>of</strong> an instrument).<br />

A stop is a baffle or o<strong>the</strong>r construction (e.g., <strong>the</strong> edge <strong>of</strong> a mirror) that limits <strong>the</strong> bundle<br />

<strong>of</strong> <strong>light</strong> that can pass through it. Two examples are given below, but stops can be used for<br />

o<strong>the</strong>r purposes such as to eliminate any stray <strong>light</strong> from <strong>the</strong> area <strong>of</strong> <strong>the</strong> beam blocked by<br />

<strong>the</strong> secondary mirror and o<strong>the</strong>r structures in front <strong>of</strong> <strong>the</strong> primary mirror.<br />

The aperture stop limits <strong>the</strong> diameter <strong>of</strong> <strong>the</strong> incoming ray bundle from <strong>the</strong> object. For a<br />

<strong>telescope</strong>, it is typically <strong>the</strong> edge <strong>of</strong> <strong>the</strong> primary mirror.<br />

A field stop limits <strong>the</strong> range <strong>of</strong> angles a <strong>telescope</strong> can accept, that is, it is <strong>the</strong> limit on <strong>the</strong><br />

<strong>telescope</strong> field <strong>of</strong> view.<br />

A pupil is an image <strong>of</strong> a stop. An entrance pupil is an image <strong>of</strong> <strong>the</strong> aperture stop formed<br />

by optics ahead <strong>of</strong> <strong>the</strong> stop (not typically encountered with <strong>telescope</strong>s, but can be used<br />

with o<strong>the</strong>r types <strong>of</strong> optics such as camera lenses or instruments mounted on a <strong>telescope</strong>).<br />

An exit pupil is an image <strong>of</strong> <strong>the</strong> aperture stop formed by optics behind <strong>the</strong> stop. For<br />

example, <strong>the</strong> secondary mirror <strong>of</strong> a Cassegrain (or Gregorian) <strong>telescope</strong> forms an image<br />

<strong>of</strong> <strong>the</strong> aperture stop and hence determines <strong>the</strong> exit pupil <strong>of</strong> <strong>the</strong> <strong>telescope</strong>. The exit pupil is<br />

<strong>of</strong>ten re-imaged within an instrument because it has unique properties for performing a<br />

number <strong>of</strong> optical functions.<br />

2.3.2. MTF/OTF<br />

The image formed by a <strong>telescope</strong><br />

departs from <strong>the</strong> ideal for<br />

diffraction by a circular aperture<br />

for many reasons, such as: 1.)<br />

non-circular primary mirrors; 2.)<br />

aberrations in <strong>the</strong> <strong>telescope</strong><br />

optics; and 3.) manufacturing<br />

errors. In addition, we need a<br />

convenient way to understand <strong>the</strong><br />

combined imaging properties <strong>of</strong><br />

<strong>the</strong> <strong>telescope</strong> plus instrumentation<br />

being used with it.<br />

The modulation transfer function,<br />

or MTF makes use <strong>of</strong> <strong>the</strong><br />

properties <strong>of</strong> Fourier transforms<br />

to provide <strong>the</strong> desired general<br />

description <strong>of</strong> <strong>the</strong> imaging<br />

Figure 2.11. Modulation <strong>of</strong> a signal – (a) input and (b) output.


properties <strong>of</strong> an optical system. It falls short <strong>of</strong> a complete description <strong>of</strong> <strong>the</strong>se properties<br />

because it does not include phase information; <strong>the</strong> optical transfer function (OTF) is <strong>the</strong><br />

MTF times <strong>the</strong> phase transfer function. However, <strong>the</strong> MTF is more convenient to<br />

manipulate and is an adequate description for a majority <strong>of</strong> applications.<br />

To illustrate <strong>the</strong> meaning <strong>of</strong> <strong>the</strong> MTF, imagine that an array <strong>of</strong> detectors is exposed to a<br />

sinusoidal input signal <strong>of</strong> period P and amplitude F(x),<br />

where f = 1/P is <strong>the</strong> spatial frequency, x is <strong>the</strong> distance along one axis <strong>of</strong> <strong>the</strong> array, a 0 is<br />

<strong>the</strong> mean height (above zero) <strong>of</strong> <strong>the</strong> pattern, and a 1 is its amplitude. These terms are<br />

indicated in Figure 2.11a. The modulation <strong>of</strong> this signal is defined as<br />

where F max and F min are <strong>the</strong> maximum and minimum values <strong>of</strong> F(x). We now imagine<br />

that we use a <strong>telescope</strong> to image a nominally identical sinusoidal pattern onto <strong>the</strong> detector<br />

array. The resulting output can be represented by<br />

where x and f are <strong>the</strong> same as in equation above, and b 0 and b 1 (f) are analogous to a 0 and<br />

a 1 (Figure 2.11b). The <strong>telescope</strong><br />

will have limited ability to image<br />

very fine details in <strong>the</strong> sinusoidal<br />

pattern. In terms <strong>of</strong> <strong>the</strong> Fourier<br />

transform <strong>of</strong> <strong>the</strong> image, this<br />

degradation <strong>of</strong> <strong>the</strong> image is<br />

described as an attenuation <strong>of</strong> <strong>the</strong><br />

high spatial frequencies in <strong>the</strong><br />

image. The dependence <strong>of</strong> <strong>the</strong><br />

signal amplitude, b 1 , on <strong>the</strong> spatial<br />

frequency, f, shows <strong>the</strong> drop in<br />

response as <strong>the</strong> spatial frequency<br />

grows (i.e., <strong>the</strong> loss <strong>of</strong> fine detail in<br />

<strong>the</strong> image). The modulation in <strong>the</strong><br />

image will be<br />

The modulation transfer factor is<br />

Figure 2.12. The MTF (schematically) corresponding to<br />

<strong>the</strong> function in Figure 2.11b.<br />

A separate value <strong>of</strong> <strong>the</strong> MT will apply at each spatial frequency; Figure 2.11a illustrates<br />

an input signal that contains a range <strong>of</strong> spatial frequencies, and Figure 2.11b shows a<br />

corresponding output in which <strong>the</strong> modulation decreases with increasing spatial<br />

frequency. This frequency dependence <strong>of</strong> <strong>the</strong> MT is expressed in <strong>the</strong> modulation transfer


function (MTF). Figure 2.12 shows <strong>the</strong> MTF corresponding to <strong>the</strong> response <strong>of</strong> Figure<br />

2.11b.<br />

In principle, <strong>the</strong> MTF provides a virtually complete specification <strong>of</strong> <strong>the</strong> imaging<br />

properties <strong>of</strong> an optical system. Computationally, <strong>the</strong> MTF can be determined by taking<br />

<strong>the</strong> absolute value <strong>of</strong> <strong>the</strong> Fourier transform, F(u), <strong>of</strong> <strong>the</strong> image <strong>of</strong> a perfect point source<br />

(this image is called <strong>the</strong> point spread function). Fourier transformation is <strong>the</strong> general<br />

ma<strong>the</strong>matical technique used to determine <strong>the</strong> frequency components <strong>of</strong> a function f(x)<br />

(see, for example, Press et al. 1986; Bracewell 2000). F(u) is defined as<br />

with inverse<br />

Only a relatively small number <strong>of</strong> functions have Fourier transforms that are easy to<br />

manipulate. With <strong>the</strong> use <strong>of</strong> computers, however, Fourier transformation is a powerful<br />

and very general technique. The Fourier transform can be generalized in a<br />

straightforward way to two dimensions, but for <strong>the</strong> sake <strong>of</strong> simplicity we will not do so<br />

here. The absolute value <strong>of</strong> <strong>the</strong> transform is<br />

where F*(u) is <strong>the</strong> complex conjugate <strong>of</strong> F(u); it is obtained by reversing <strong>the</strong> sign <strong>of</strong> all<br />

imaginary terms in F(u).<br />

The point spread function is<br />

<strong>the</strong> image a system forms<br />

<strong>of</strong> a perfectly sharp point<br />

input image. If f(x)<br />

represents <strong>the</strong> point spread<br />

function, |F(u)|/|F(0)| is <strong>the</strong><br />

MTF <strong>of</strong> <strong>the</strong> system, with u<br />

<strong>the</strong> spatial frequency. This<br />

formulation holds because a<br />

sharp impulse, represented<br />

ma<strong>the</strong>matically by a <br />

function, contains all<br />

frequencies equally (that is,<br />

its Fourier transform is a<br />

constant). Hence <strong>the</strong><br />

Fourier transform <strong>of</strong> <strong>the</strong><br />

image formed from an input<br />

sharp impulse gives <strong>the</strong><br />

spatial frequency response<br />

<strong>of</strong> <strong>the</strong> optical system.<br />

Figure 2.13. MTF <strong>of</strong> a round <strong>telescope</strong> with no central obscuration.<br />

Spatial frequencies are in units <strong>of</strong> D/λ.


The MTF is normalized to unity at spatial frequency 0 by this definition. As emphasized<br />

in Figure 2.12, <strong>the</strong> response at zero frequency cannot be measured directly but must be<br />

extrapolated from higher frequencies.<br />

The image <strong>of</strong> an entire linear optical system is <strong>the</strong> convolution <strong>of</strong> <strong>the</strong> images from each<br />

element. By <strong>the</strong> "convolution <strong>the</strong>orem", its MTF can be determined by multiplying<br />

toge<strong>the</strong>r <strong>the</strong> MTFs <strong>of</strong> its constituent elements, and <strong>the</strong> resulting image is determined by<br />

inverse transforming <strong>the</strong> resulting MTF. The multiplication occurs on a frequency by<br />

frequency basis, that is, if <strong>the</strong> first system has MTF 1 (f) and <strong>the</strong> second MTF 2 (f), <strong>the</strong><br />

combined system has MTF(f) = MTF 1 (f) MTF 2 (f). The overall resolution capability <strong>of</strong><br />

complex optical systems can be more easily determined in this way than by brute force<br />

image convolution.<br />

We now show <strong>the</strong> power <strong>of</strong> <strong>the</strong> MTF by illustrating <strong>the</strong> calculation <strong>of</strong> <strong>the</strong> Airy function.<br />

We define <strong>the</strong> box function in two dimensions as<br />

Thus, we describe a round aperture <strong>of</strong> radius r 0 functionally as<br />

The corresponding MTF is<br />

<br />

(see Figure 2.13). The intensity in <strong>the</strong> image is<br />

<br />

<br />

This result is identical to equation (2.1) describing <strong>the</strong> diffraction-limited image.<br />

2.4. Telescope Optimization<br />

2.4.1. Wide Field<br />

The classic <strong>telescope</strong> designs all rely<br />

on <strong>the</strong> imaging properties <strong>of</strong> a<br />

paraboloid, with secondary mirrors to<br />

relay <strong>the</strong> image formed by <strong>the</strong> primary<br />

mirror to a second focus. What would<br />

happen if no image were demanded <strong>of</strong><br />

<strong>the</strong> primary, but <strong>the</strong> conic sections on<br />

<strong>the</strong> primary and secondary mirrors<br />

were selected to provide optimum<br />

imaging as a unit One answer to this<br />

Figure 2.14. Three-mirror wide field <strong>telescope</strong> design<br />

after Baker and Paul.<br />

question is <strong>the</strong> Ritchey-Cretién <strong>telescope</strong>. In this design, a suitable selection <strong>of</strong> hyperbolic<br />

primary and secondary toge<strong>the</strong>r can compensate for <strong>the</strong> spherical and comatic aberrations


(to third order in an expansion in angle <strong>of</strong> incidence) and provide large fields, up to an<br />

outer boundary where astigmatism becomes objectionable. An alternative, <strong>the</strong> aplanatic<br />

Gregorian <strong>telescope</strong>, has ellipsoidal primary and secondary, and similar wide-field<br />

performance.<br />

Still larger fields can be achieved by adding optics specifically to compensate for <strong>the</strong><br />

aberrations. A classic example is <strong>the</strong> Schmidt camera, which has a full-aperture refractive<br />

corrector in front <strong>of</strong> its spherical primary mirror. The corrector imposes spherical<br />

aberration on <strong>the</strong> input beam in <strong>the</strong> same amount but opposite sign as <strong>the</strong> aberration <strong>of</strong><br />

<strong>the</strong> primary, producing well-corrected images over a large field. The Maksutov and<br />

Schmidt-Cassegrain designs achieve similar ends with correctors <strong>of</strong> different types.<br />

However, refractive elements larger than 1 – 1.5m in diameter become prohibitively<br />

difficult to mount and maintain in figure. A variety <strong>of</strong> designs for correctors near <strong>the</strong><br />

focal plane are also possible (e.g., Epps & Fabricant 1997). A simple version invented by<br />

F. E. Ross uses two lenses, one positive and <strong>the</strong> o<strong>the</strong>r negative to cancel each o<strong>the</strong>r in<br />

optical power but to correct coma. Wynne (1974) proposed a more powerful device with<br />

three lenses, <strong>of</strong> positive-negative-positive power and all with spherical surfaces. A<br />

Wynne corrector can correct for chromatic aberration, field curvature, spherical<br />

aberration, coma, and astigmatism. Such approaches are used with all <strong>the</strong> wide field<br />

cameras on 10-m-class <strong>telescope</strong>s. These optical trains not only correct <strong>the</strong> field but also<br />

can provide for spectrally active components, such as compensators for <strong>the</strong> dispersion <strong>of</strong><br />

<strong>the</strong> atmosphere that allow high quality images to be obtained at <strong>of</strong>f-zenith pointings and<br />

over broad spectral bands (e.g., Fabricant et al. 2004).<br />

Even wider fields can be obtained by adding large reflective elements to <strong>the</strong> optical train<br />

<strong>of</strong> <strong>the</strong> <strong>telescope</strong>, an approach pioneered by Baker and Paul, see Figure 2.14. The basic<br />

concept is that, by making <strong>the</strong> second mirror spherical, <strong>the</strong> <strong>light</strong> is deviated in a similar<br />

way as it would be by <strong>the</strong> corrector lens <strong>of</strong> a Schmidt camera. The third, spherical, mirror<br />

brings <strong>the</strong> <strong>light</strong> to <strong>the</strong> <strong>telescope</strong> focal plane over a large field. The Baker-Paul concept is a<br />

very powerful one and underlies <strong>the</strong> large field <strong>of</strong> <strong>the</strong> Large Synoptic Survey Telescope<br />

(LSST). . With three or more mirror surfaces to play with, optical designers can achieve a<br />

high degree <strong>of</strong> image correction in o<strong>the</strong>r ways also. For example, a three-mirror<br />

anastigmat is corrected simultaneously for spherical aberration, coma, and astigmatism.<br />

2.4.2. Infrared<br />

The basic optical considerations for groundbased <strong>telescope</strong>s apply in <strong>the</strong> infrared as well.<br />

However, in <strong>the</strong> <strong>the</strong>rmal infrared (wavelengths longer than about 2m), <strong>the</strong> emission <strong>of</strong><br />

<strong>the</strong> <strong>telescope</strong> is <strong>the</strong> dominant signal on <strong>the</strong> detectors and needs to be minimized to<br />

maximize <strong>the</strong> achievable signal to noise on astronomical sources. It is not possible to cool<br />

any <strong>of</strong> <strong>the</strong> exposed parts <strong>of</strong> <strong>the</strong> <strong>telescope</strong> because <strong>of</strong> atmospheric condensation. However,<br />

<strong>the</strong> sky is both colder and <strong>of</strong> lower emissivity (since one looks in bands where it is<br />

transparent) than <strong>the</strong> structure <strong>of</strong> <strong>the</strong> <strong>telescope</strong>. Therefore, one minimizes <strong>the</strong> structure<br />

visible to <strong>the</strong> detector by using a modest sized secondary mirror (e.g., f/15), matching it<br />

with a small central hole in <strong>the</strong> primary mirror, keeping <strong>the</strong> secondary supports and o<strong>the</strong>r


structures in <strong>the</strong> beam with thin cross-sections, and most importantly, removing <strong>the</strong><br />

baffles. The instruments operate at cryogenic temperatures; by forming a pupil within<br />

<strong>the</strong>ir optical trains and putting a tight stop at it, <strong>the</strong> view <strong>of</strong> warm structures can be<br />

minimized without losing <strong>light</strong> from <strong>the</strong> astronomical sources. However, <strong>the</strong> pupil image<br />

generally is not perfectly sharp and <strong>the</strong> stop is <strong>the</strong>refore made a bit oversized. To avoid<br />

excess emission from <strong>the</strong> area around <strong>the</strong> primary mirror entering <strong>the</strong> system, <strong>of</strong>ten <strong>the</strong><br />

secondary mirror is made s<strong>light</strong>ly undersized, so its edge becomes <strong>the</strong> aperture stop <strong>of</strong> <strong>the</strong><br />

<strong>telescope</strong>. Only emission from <strong>the</strong> sky enters <strong>the</strong> instrument over <strong>the</strong> edge <strong>of</strong> <strong>the</strong><br />

secondary, and in high-quality infrared atmospheric windows <strong>the</strong> sky is more than an<br />

order <strong>of</strong> magnitude less emissive than <strong>the</strong> primary mirror cell would be.<br />

The emissivity <strong>of</strong> <strong>the</strong> <strong>telescope</strong> mirrors needs to be kept low. If <strong>the</strong> <strong>telescope</strong> uses<br />

conventional aluminum coatings, <strong>the</strong>y must be kept very clean. Better performance can<br />

be obtained with silver or gold coatings. Obviously, it is also important to avoid<br />

additional warm mirrors in <strong>the</strong> optical train o<strong>the</strong>r than <strong>the</strong> primary and secondary.<br />

Therefore, <strong>the</strong> secondary mirror is called upon to perform a number <strong>of</strong> functions besides<br />

helping form an image. Because <strong>of</strong> <strong>the</strong> variability <strong>of</strong> <strong>the</strong> atmospheric emission, it is<br />

desirable to compare <strong>the</strong> image containing <strong>the</strong> source with an image <strong>of</strong> adjacent sky<br />

rapidly, and an articulated secondary can do this in a fraction <strong>of</strong> a second if desired.<br />

Modulating <strong>the</strong> signal with <strong>the</strong> secondary mirror has <strong>the</strong> important advantage that <strong>the</strong><br />

beam passes through nearly <strong>the</strong> identical column <strong>of</strong> air as it traverses <strong>the</strong> <strong>telescope</strong> (and<br />

even from high into <strong>the</strong> atmosphere), so any fluctuations in <strong>the</strong> infrared emission along<br />

this column are common to both mirror positions. A suitably designed secondary mirror<br />

can also improve <strong>the</strong> imaging, ei<strong>the</strong>r by compensating for image motion or even<br />

correcting <strong>the</strong> wavefront errors imposed by <strong>the</strong> atmosphere (see Chapter 7).<br />

The ultimate solution to <strong>the</strong>se issues is to place <strong>the</strong> infrared <strong>telescope</strong> in space, where it<br />

can be cooled sufficiently that it no longer dominates <strong>the</strong> signals. Throughout <strong>the</strong> <strong>the</strong>rmal<br />

infrared, <strong>the</strong> foreground signals can be reduced by a factor <strong>of</strong> a million or more in space<br />

compared with <strong>the</strong> ground, and hence phenomenally greater sensitivity can be achieved.<br />

Obviously, space also <strong>of</strong>fers <strong>the</strong> advantage <strong>of</strong> being above <strong>the</strong> atmosphere and hence free<br />

<strong>of</strong> atmospheric absorption, so <strong>the</strong> entire infrared range is accessible for observation. The<br />

price <strong>of</strong> operation in space includes not only cost, but restrictions on <strong>the</strong> maximum<br />

aperture (being mitigated with JWST), and <strong>the</strong> inability to change instruments or even to<br />

include all <strong>the</strong> instrument types desired. None<strong>the</strong>less, astronomy has benefited from a<br />

series <strong>of</strong> very successful cryogenic <strong>telescope</strong>s in space – <strong>the</strong> Infrared Astronomy Satellite<br />

(IRAS), <strong>the</strong> Infrared Space Observatory (ISO), <strong>the</strong> Spitzer Telescope, Akari, <strong>the</strong> Cosmic<br />

Background Explorer (COBE), and <strong>the</strong> Wide-Field Infrared Survey Explorer (WISE), to<br />

name some examples.<br />

2.4.3. Radio<br />

Radio <strong>telescope</strong>s are typically <strong>of</strong> conventional parabolic-primary-mirror, prime-focus<br />

design. The apertures are very large, so <strong>the</strong> blockage <strong>of</strong> <strong>the</strong> beam by <strong>the</strong> receivers at <strong>the</strong><br />

prime focus is insignificant. The primary mirrors have short focal lengths, f-ratios ~ 0.5,<br />

to keep <strong>the</strong> <strong>telescope</strong> compact and help provide a rigid structure. Sometimes <strong>the</strong> entire


<strong>telescope</strong> is designed in a way that <strong>the</strong> flexure as it is pointed in different directions<br />

occurs in a way that preserves <strong>the</strong> figure <strong>of</strong> <strong>the</strong> primary – <strong>the</strong>se designs are described as<br />

deforming homologously. For example, <strong>the</strong> 100-m aperture Effelsberg Telescope flexes<br />

by up to 6cm as it is pointed to different elevations, but maintains its paraboloidal figure<br />

to an accuracy <strong>of</strong> ~ 4mm.<br />

Telescopes for <strong>the</strong> mm- and sub-mm wave regimes are generally smaller and usually are<br />

built in a Cassegrain configuration, <strong>of</strong>ten with secondary mirrors that can be chopped or<br />

nutated over small angles to help<br />

compensate for background<br />

emission. They can be<br />

considered to be intermediate in<br />

design between longerwavelength<br />

radio <strong>telescope</strong>s and<br />

infrared <strong>telescope</strong>s. Ano<strong>the</strong>r<br />

exception to <strong>the</strong> general design<br />

trend is <strong>the</strong> Arecibo <strong>telescope</strong>,<br />

which has a spherical primary to<br />

allow steering <strong>the</strong> beam in<br />

different directions while <strong>the</strong><br />

primary is fixed. The spherical<br />

aberration is corrected by optics<br />

near <strong>the</strong> prime focus <strong>of</strong> <strong>the</strong><br />

<strong>telescope</strong> in this case.<br />

Radio receivers are uniformly <strong>of</strong><br />

coherent detector design; only in<br />

<strong>the</strong> mm- and sub-mm realms are<br />

bolometers used for continuum<br />

measurements and lowresolution<br />

spectroscopy. The<br />

operation <strong>of</strong> <strong>the</strong>se devices will Figure 2.15. The Byrd Green Bank Telescope.<br />

be described in more detail in<br />

Chapter 8. For this discussion, we need to understand that all such receivers are limited<br />

by <strong>the</strong> antenna <strong>the</strong>orem, which states that <strong>the</strong>y are sensitive only to <strong>the</strong> central peak <strong>of</strong> <strong>the</strong><br />

diffraction pattern <strong>of</strong> <strong>the</strong> <strong>telescope</strong>, and to only a single polarization.<br />

This behavior modifies how <strong>the</strong> imaging properties <strong>of</strong> <strong>the</strong> <strong>telescope</strong> are described. The<br />

primary measure <strong>of</strong> <strong>the</strong> quality <strong>of</strong> <strong>the</strong> <strong>telescope</strong> optics is beam efficiency, <strong>the</strong> ratio <strong>of</strong> <strong>the</strong><br />

power from a point source in <strong>the</strong> central peak <strong>of</strong> <strong>the</strong> image to <strong>the</strong> power in <strong>the</strong> entire<br />

image. The diffraction rings <strong>of</strong> <strong>the</strong> Airy pattern appear to radio astronomers as potential<br />

regions <strong>of</strong> unwanted sensitivity to sources away from <strong>the</strong> one at which <strong>the</strong> <strong>telescope</strong> is<br />

pointed. This issue is significant because <strong>of</strong> <strong>the</strong> relatively low angular resolution and high<br />

sensitivity <strong>of</strong> single-dish radio <strong>telescope</strong>s (as opposed to interferometers, see Chapter 9).<br />

The sensitive regions <strong>of</strong> <strong>the</strong> <strong>telescope</strong> beam outside <strong>the</strong> main Airy peak are called<br />

sidelobes. Additional sidelobes are produced by imperfections in <strong>the</strong> mirror surface. A


consideration in receiver design is <strong>the</strong> response over <strong>the</strong> primary mirror, called <strong>the</strong><br />

illumination pattern, to minimize <strong>the</strong> sidelobes from <strong>the</strong> image. Sidelobes also arise<br />

through diffraction from structures in <strong>the</strong> beam, particularly <strong>the</strong> supports for <strong>the</strong> primefocus<br />

instruments (or secondary mirror). The Greenbank Telescope (Figure 2.15)<br />

eliminates <strong>the</strong>m with an <strong>of</strong>f-axis primary that brings <strong>the</strong> focus to <strong>the</strong> side <strong>of</strong> <strong>the</strong> incoming<br />

beam.<br />

2.4.4. X-ray<br />

Certain materials have indices <strong>of</strong><br />

refraction in <strong>the</strong> 0.1 – 10 kev range<br />

that are s<strong>light</strong>ly less than 1 (by ~<br />

0.01 at low energies and only ~<br />

0.0001 at high). Consequently, at<br />

grazing angles <strong>the</strong>se materials<br />

reflect X-rays by total external<br />

reflection; however, at <strong>the</strong> high<br />

energy end, <strong>the</strong> angle <strong>of</strong> incidence<br />

can be only <strong>of</strong> order 1 o . In addition,<br />

<strong>the</strong> images formed by grazing<br />

incidence <strong>of</strong>f a paraboloid have<br />

severe astigmatism <strong>of</strong>f-axis, so two<br />

reflections are needed for good<br />

images, one <strong>of</strong>f a paraboloid and<br />

<strong>the</strong> o<strong>the</strong>r <strong>of</strong>f a hyperboloid or<br />

Figure 2.16. A Wolter Type I grazing incidence X-ray<br />

<strong>telescope</strong> – vertical scale exaggerated for clarity.<br />

ellipsoid. Given <strong>the</strong>se constraints, <strong>the</strong> optical arrangements <strong>of</strong> <strong>telescope</strong>s based on this<br />

type <strong>of</strong> reflection look relatively familiar (Figure 2.16) – for example, a reflection <strong>of</strong>f a<br />

paraboloid followed by one <strong>of</strong>f a hyperboloid for a Wolter Type-1 geometry (Figure<br />

2.16). A Wolter Type-2 <strong>telescope</strong> uses a convex hyperboloid for <strong>the</strong> second reflection to<br />

give a longer focal length and larger image scale, while a Wolter Type-3 <strong>telescope</strong> uses a<br />

convex paraboloid and concave ellipsoid. The on-axis imaging quality <strong>of</strong> such <strong>telescope</strong>s<br />

is strongly dependent on <strong>the</strong> quality <strong>of</strong> <strong>the</strong> reflecting surfaces. The constraints in optical<br />

design already imposed by <strong>the</strong> grazing incidence reflection make it difficult to correct <strong>the</strong><br />

optics well for large fields and <strong>the</strong> imaging quality degrades significantly for fields larger<br />

than a few arcminutes in radius. The areal efficiency <strong>of</strong> <strong>the</strong>se optical trains is low, since<br />

<strong>the</strong>ir mirrors collect photons only over a narrow annulus. It is <strong>the</strong>refore common to nest a<br />

number <strong>of</strong> optical trains to increase <strong>the</strong> collection <strong>of</strong> photons without increasing <strong>the</strong> total<br />

size <strong>of</strong> <strong>the</strong> assembled <strong>telescope</strong>.


As an example, we consider Chandra (Figure 2.17). Its Wolter Type-1 <strong>telescope</strong> has a<br />

diameter <strong>of</strong> 1.2m, within which <strong>the</strong>re are four nested optical trains, which toge<strong>the</strong>r<br />

provide a total collecting area <strong>of</strong> 1100 cm 2 (i.e., ~ 10% <strong>of</strong> <strong>the</strong> total entrance aperture).<br />

Figure 2.17. The Chandra <strong>telescope</strong>.<br />

The focal length is 10m, <strong>the</strong> angles <strong>of</strong><br />

incidence onto <strong>the</strong> mirror surfaces range<br />

from 27 to 51 arcmin, and <strong>the</strong> reflection<br />

material is iridium. The <strong>telescope</strong><br />

efficiency is reasonably good from 0.1 to 7<br />

keV. The Chandra design emphasizes<br />

angular resolution. The on-axis images are<br />

0.5 arcsec in diameter (Figure 2.18) but<br />

degrade by more than an order <strong>of</strong><br />

magnitude at an <strong>of</strong>f-axis radius <strong>of</strong> 10<br />

arcmin (Figure 2.19).<br />

The Chandra design can be compared with<br />

that <strong>of</strong> XMM-Newton, which emphasizes<br />

collecting area. The latter <strong>telescope</strong> has<br />

three modules <strong>of</strong> 58 nested optical trains,<br />

Figure 2.18. On-axis image quality for Chandra.<br />

Half <strong>of</strong> <strong>the</strong> energy is contained within an image<br />

radius <strong>of</strong> 0.418 arcsec.<br />

each <strong>of</strong> diameter 70cm and with a collecting area <strong>of</strong> 2000cm 2 (~ 50% <strong>of</strong> <strong>the</strong> total entrance<br />

aperture). The total collecting area is 6000cm 2 , a factor <strong>of</strong> five greater than for Chandra.<br />

The range <strong>of</strong> grazing angles, 18 – 40 arcmin, is smaller than for Chandra resulting in<br />

greater high energy (~ 10 keV) efficiency. The mirrors are not as high in quality as <strong>the</strong><br />

Chandra ones, so <strong>the</strong> on-axis images are an order <strong>of</strong> magnitude larger in diameter.


For reflecting surfaces,<br />

Chandra and XMM-Newton<br />

used single materials (e.g.,<br />

platinum, iridium, gold) in<br />

grazing incidence. This<br />

approach drops rapidly in<br />

efficiency (or equivalently<br />

works only at more and more<br />

grazing incidence) with<br />

increased energy and is not<br />

effective above about 10 kev.<br />

Higher energy photons can be<br />

reflected using <strong>the</strong> Bragg<br />

effect. Figure 2.20 shows how<br />

a crystal lattice can reflect by<br />

constructive interference.<br />

<br />

With identical layers, <strong>the</strong><br />

reflection would be over a<br />

restricted energy range. By<br />

grading multiple layers, broad<br />

Figure 2.19. The Chandra images degrade rapidly with<br />

increasing <strong>of</strong>f-axis angle.<br />

ranges are reflected. The top layers have large spacing (d) and reflect <strong>the</strong> low energies,<br />

while lower layers have smaller d for higher energies. A stack <strong>of</strong> up to 200 layers yields<br />

an efficient reflector.<br />

NuStar(to be launched in<br />

2011) uses this approach to<br />

make reflecting optics<br />

working up to 79eV. It will use<br />

a standard Wolter design with<br />

multilayer surfaces deposited<br />

on thin glass substrates,<br />

slumped to <strong>the</strong> correct shape.<br />

It will have 130 reflecting<br />

shells, with a focal length <strong>of</strong><br />

10 meters. The <strong>telescope</strong> will<br />

be deployed after launch so<br />

<strong>the</strong> spacecraft fits within <strong>the</strong><br />

launch constraints.<br />

Figure 2.20. Bragg Effect. The dots represent <strong>the</strong> atoms in a<br />

regular crystal lattice. Constructive interference occurs at<br />

specific grazing incidence reflection angles.<br />

2.5.Modern Optical-Infrared Telescopes<br />

2.5.1. 10-meter-class Telescopes


For many years, <strong>the</strong> Palomar 5-m <strong>telescope</strong> was considered <strong>the</strong> ultimate large<br />

groundbased <strong>telescope</strong>; flexure in <strong>the</strong> primary mirror was thought to be a serious obstacle<br />

to construction <strong>of</strong> larger ones. The benefits from larger <strong>telescope</strong>s were also argued to be<br />

modest. If <strong>the</strong> image size remains <strong>the</strong> same (e.g., is set by a constant level <strong>of</strong> seeing), <strong>the</strong>n<br />

<strong>the</strong> gain in sensitivity with a background limited detector goes only as <strong>the</strong> diameter <strong>of</strong> <strong>the</strong><br />

<strong>telescope</strong> primary mirror.<br />

This situation changed with dual advances. The size limit implied by <strong>the</strong> 5-m primary<br />

mirror can be violated by application <strong>of</strong> a variety <strong>of</strong> techniques to hold <strong>the</strong> mirror figure<br />

in <strong>the</strong> face <strong>of</strong> flexure. The images produced by <strong>the</strong> <strong>telescope</strong> can be analyzed to<br />

determine exactly what adjustments are needed to its primary to apply <strong>the</strong>se techniques<br />

and fix any issues with flexure. As a result, <strong>the</strong> primary can be made both larger than 5<br />

meters in diameter and <strong>of</strong> substantially lower mass, since <strong>the</strong> rigidity can be provided by<br />

external controls ra<strong>the</strong>r than intrinsic stiffness. In addition, it was realized that much <strong>of</strong><br />

<strong>the</strong> degradation <strong>of</strong> images due to seeing was occurring within <strong>the</strong> <strong>telescope</strong> dome, due to<br />

air currents arising from warm surfaces. The most significant examples were within <strong>the</strong><br />

<strong>telescope</strong> itself – <strong>the</strong> massive primary mirror and <strong>the</strong> correspondingly massive steel<br />

structures required to support it and point it accurately. By reducing <strong>the</strong> mass <strong>of</strong> <strong>the</strong><br />

primary with modern<br />

control methods, <strong>the</strong><br />

entire <strong>telescope</strong> could be<br />

made less massive,<br />

resulting in a<br />

corresponding reduction<br />

in its heat capacity and a<br />

faster approach to<br />

<strong>the</strong>rmal equilibrium with<br />

<strong>the</strong> ambient temperature.<br />

This adjustment is fur<strong>the</strong>r<br />

hastened by aggressive<br />

ventilation <strong>of</strong> <strong>the</strong><br />

<strong>telescope</strong> enclosure,<br />

including providing it<br />

with large vents that<br />

almost place <strong>the</strong><br />

<strong>telescope</strong> in <strong>the</strong> open air<br />

while observing. The<br />

Figure 2.21. Keck Telescope<br />

gains with <strong>the</strong> current generation <strong>of</strong> large <strong>telescope</strong>s <strong>the</strong>refore derive from both <strong>the</strong><br />

increase in collecting area and <strong>the</strong> reduction in image size. A central feature <strong>of</strong> plans for<br />

even larger <strong>telescope</strong>s is sophisticated adaptive optics systems to shrink <strong>the</strong>ir image sizes<br />

fur<strong>the</strong>r (see Chapter 7).


Three basic approaches have been developed for large groundbased <strong>telescope</strong>s. The Keck<br />

Telescopes (Figure 2.21), Gran Telescopio Canarias (GTC), Hobby-Everly Telescope<br />

(HET), and South African Large Telescope (SALT) exemplify <strong>the</strong> use <strong>of</strong> a segmented<br />

primary mirror. The 10-m primary mirror <strong>of</strong> Keck is an array <strong>of</strong> 36 hexagonal mirrors,<br />

each 0.9-m on a side, and made <strong>of</strong> Zerodur glass-ceramic (this material has relatively low<br />

<strong>the</strong>rmal change with<br />

temperature). The<br />

positions <strong>of</strong> <strong>the</strong>se<br />

segments relative to<br />

each o<strong>the</strong>r are sensed<br />

by capacitive sensors.<br />

A specialized alignment<br />

camera is used to set<br />

<strong>the</strong> segments in tip and<br />

tilt and <strong>the</strong>n <strong>the</strong> mirror<br />

is locked under control<br />

<strong>of</strong> <strong>the</strong> edge sensors. The<br />

mirrors must be<br />

adjusted very<br />

accurately in piston for<br />

<strong>the</strong> <strong>telescope</strong> to operate<br />

in a diffraction-limited<br />

mode. The alignment<br />

camera allows for<br />

adjustment <strong>of</strong> each pair<br />

<strong>of</strong> segments in this<br />

coordinate by<br />

Figure 2.22. The MMT.<br />

interfering <strong>the</strong> <strong>light</strong> in a small aperture that straddles <strong>the</strong> edges <strong>of</strong> <strong>the</strong> segments. As <strong>the</strong><br />

first <strong>of</strong> <strong>the</strong> current generation <strong>of</strong> large <strong>telescope</strong>s, <strong>the</strong> Keck Observatory has a broad range<br />

<strong>of</strong> instrumentation for <strong>the</strong> optical and near infrared, as well as a high-performance<br />

adaptive optics system.<br />

The VLT, Subaru, and Gemini <strong>telescope</strong>s use a thin monolithic plate for <strong>the</strong> optical<br />

element <strong>of</strong> <strong>the</strong> primary mirror. We take <strong>the</strong> VLT as an example – it is actually four<br />

identical examples. They each have 8.2-meter primary mirrors <strong>of</strong> Zerodur that are only<br />

0.175 meters thick. A VLT primary is supported against flexure by 150 actuators that are<br />

controlled by image analysis at an interval <strong>of</strong> a couple <strong>of</strong> times per minute. Again, as a<br />

mature observatory, a wide range <strong>of</strong> instrumentation is available.<br />

The MMT (Figure 2.22), Magellan, and LBT Telescopes (primaries respectively 6.5, 6.5,<br />

and 8.4m in diameter) are based on a monolithic primary mirror design that is deeply<br />

relieved in <strong>the</strong> back to reduce <strong>the</strong> mass and <strong>the</strong>rmal inertia. Use <strong>of</strong> a polishing lap with a<br />

computer-controlled shape allows <strong>the</strong> manufacture <strong>of</strong> very fast mirrors; e.g., <strong>the</strong> two for<br />

<strong>the</strong> LBT are f/1.14, which allows an enclosure <strong>of</strong> minimum size. The LBT is developing<br />

pairs <strong>of</strong> instruments for <strong>the</strong> individual <strong>telescope</strong>s, such as prime focus cameras (with a<br />

red camera at <strong>the</strong> focus <strong>of</strong> one mirror and a blue camera for <strong>the</strong> o<strong>the</strong>r), or twin


spectrographs on both sides. The outputs <strong>of</strong> <strong>the</strong> two sides <strong>of</strong> <strong>the</strong> <strong>telescope</strong> can also be<br />

combined for operation as an interferometer. The design with <strong>the</strong> primary mirrors on a<br />

single mount eliminates large path length differences between <strong>the</strong>m and gives <strong>the</strong><br />

interferometric applications a uniquely large field <strong>of</strong> view compared with o<strong>the</strong>r<br />

approaches.<br />

2.4.2. Wave Front Sensing<br />

To make <strong>the</strong> adjustments that maintain <strong>the</strong>ir image quality, all <strong>of</strong> <strong>the</strong>se <strong>telescope</strong>s depend<br />

on frequent and accurate<br />

measurement <strong>of</strong> <strong>the</strong> <strong>telescope</strong><br />

aberrations that result from<br />

flexure and <strong>the</strong>rmal drift. A<br />

common way to make <strong>the</strong>se<br />

measurements is <strong>the</strong> Shack-<br />

Hartmann Sensor, which<br />

operates on <strong>the</strong> principle<br />

illustrated in Figure 2.23 (see<br />

http://www.wavefrontscience<br />

s.com/Historical%20Develop<br />

ment.pdf for a historical<br />

description). A “perfect”<br />

optical system maintains<br />

wavefronts that are plane or<br />

spherical. In <strong>the</strong> Shack-<br />

Hartmann Sensor, <strong>the</strong><br />

wavefront is divided at a<br />

pupil by an array <strong>of</strong> small Figure 2.23. Principle <strong>of</strong> operation <strong>of</strong> a Shack-Hartmann Sensor.<br />

lenslets. In Figure 2.23, <strong>the</strong><br />

situation for a perfect plane wavefront is shown in dashed lines going in to <strong>the</strong> lenslet<br />

array. Each lenslet images its piece <strong>of</strong> <strong>the</strong> wavefront onto <strong>the</strong> CCD (<strong>the</strong> imaging process<br />

is shown as <strong>the</strong> paths <strong>of</strong> <strong>the</strong> outer rays, not <strong>the</strong> wavefronts) directly behind <strong>the</strong> lens, on its<br />

optical axis. For <strong>the</strong> plane input wavefront, <strong>the</strong>se images will form a grid that is<br />

uniformly spaced.<br />

Aberrations impose deviations on <strong>the</strong> wavefronts. An example is shown as a solid line in<br />

Figure 2.23. Each lenslet will see a locally tilted portion <strong>of</strong> <strong>the</strong> incoming wavefront, as if<br />

it saw an unaberrated wavefront from <strong>the</strong> wrong direction. As a result, <strong>the</strong> images from<br />

<strong>the</strong> individual lenslets will be tilted relative to <strong>the</strong> optical axis <strong>of</strong> <strong>the</strong> lens and displaced<br />

when <strong>the</strong>y reach <strong>the</strong> CCD. A simple measurement <strong>of</strong> <strong>the</strong> positions <strong>of</strong> <strong>the</strong>se images can<br />

<strong>the</strong>n be used to calculate <strong>the</strong> shape <strong>of</strong> <strong>the</strong> incoming wavefront, and hence to determine <strong>the</strong><br />

aberrations in <strong>the</strong> optics from which it was delivered. These can be corrected by a<br />

combination <strong>of</strong> adjustments on <strong>the</strong> primary mirror and motions <strong>of</strong> <strong>the</strong> secondary (e.g., <strong>the</strong><br />

latter to correct focus changes).<br />

2.4.3. Telescopes <strong>of</strong> <strong>the</strong> Future


The methods developed for control <strong>of</strong> <strong>the</strong> figure <strong>of</strong> large primary mirrors on <strong>the</strong> ground<br />

have been adopted for <strong>the</strong> James Webb Space Telescope, in this case so <strong>the</strong> 6.5-m<br />

primary mirror can be folded to fit within <strong>the</strong> shroud <strong>of</strong> <strong>the</strong> launch rocket. After launch,<br />

<strong>the</strong> primary is unfolded and <strong>the</strong>n a series <strong>of</strong> ever more demanding tests and adjustments<br />

will bring it into proper figure. The demands for very <strong>light</strong> weight have led to a primary<br />

mirror <strong>of</strong> 18 segments <strong>of</strong> beryllium, and fast optically (roughly f/1.5, but after reflection<br />

<strong>of</strong>f <strong>the</strong> convex secondary <strong>the</strong> <strong>telescope</strong> has a final f/ratio <strong>of</strong> about 17). Periodic<br />

measurements with <strong>the</strong> near infrared camera will be used to monitor <strong>the</strong> primary mirror<br />

figure and adjust it as necessary for optimum performance. The overall design is a threemirror<br />

anastigmat (meaning it is corrected fully for spherical aberration, coma, and<br />

astigmatism), with a fourth mirror for fine steering <strong>of</strong> <strong>the</strong> images.<br />

There are a number <strong>of</strong> proposals for 30-meter class groundbased <strong>telescope</strong>s. Given <strong>the</strong><br />

slow gain in sensitivity with increasing aperture for constant image diameter, all <strong>of</strong> <strong>the</strong>se<br />

proposals are based on <strong>the</strong> potential for fur<strong>the</strong>r improvements in image quality to<br />

accompany <strong>the</strong> increase in size. These gains will be achieved with multi-conjugate<br />

adaptive optics (MCAO). MCAO is based on multiple wavefront-correcting mirrors, each<br />

mirror placed in <strong>the</strong> optics to work at a particular elevation in <strong>the</strong> atmosphere (or more<br />

accurately, at a given range from <strong>the</strong> <strong>telescope</strong>). Laser beacons are directed into <strong>the</strong><br />

atmosphere and <strong>the</strong> returned signals (due, e.g., to scattering) along with those from<br />

natural guide stars are analyzed to determine <strong>the</strong> corrections to apply to <strong>the</strong>se mirrors.<br />

Such a system achieves a three-dimensional correction <strong>of</strong> <strong>the</strong> atmospheric seeing and can<br />

1.) extend <strong>the</strong> corrections to shorter wavelengths, i.e., <strong>the</strong> optical; 2.) increase <strong>the</strong> size <strong>of</strong><br />

<strong>the</strong> compensated field <strong>of</strong> view; and 3.) improve <strong>the</strong> uniformity <strong>of</strong> <strong>the</strong> images over this<br />

field. See Chapter 7 for additional discussion <strong>of</strong> MCAO.<br />

One proposal, <strong>the</strong> Thirty Meter Telescope (TMT), is from a partnership <strong>of</strong> Canada,<br />

CalTech, <strong>the</strong><br />

<strong>University</strong> <strong>of</strong><br />

California, Japan,<br />

China, and India<br />

(some still in<br />

“observer” status).<br />

This <strong>telescope</strong> would<br />

build on <strong>the</strong> Keck<br />

Telescope approach.<br />

Its primary mirror<br />

would have 492<br />

segments (up from 36<br />

for Keck). The<br />

European Sou<strong>the</strong>rn<br />

Observatory<br />

proposed to build a<br />

100-m <strong>telescope</strong><br />

called <strong>the</strong><br />

Figure 2.24. Artist’s concept <strong>of</strong> <strong>the</strong> Giant Magellan Telescope.


Overwhelming Large Telescope (OWL). It was not clear what <strong>the</strong>y would name <strong>the</strong> next<br />

generation one, so <strong>the</strong>y have scaled down to <strong>the</strong> European Extremely Large Telescope<br />

(E-ELT), a 42-m aperture segmented mirror design. The Giant Magellan Telescope<br />

(GMT, Figure 2.24) is being developed by an international consortium, including <strong>the</strong><br />

<strong>University</strong> <strong>of</strong> <strong>Arizona</strong> plus Astronomy Australia Ltd., <strong>the</strong> Australian National <strong>University</strong>,<br />

<strong>the</strong> Carnegie Institution for Science, Harvard <strong>University</strong>, <strong>the</strong> Korea Astronomy and Space<br />

Science Institute, <strong>the</strong> Smithsonian Institution, Texas A&M <strong>University</strong>, <strong>the</strong> <strong>University</strong> <strong>of</strong><br />

Texas at Austin, and <strong>the</strong> <strong>University</strong> <strong>of</strong> Chicago. It will be based on a close-packed<br />

arrangement <strong>of</strong> seven 8.4-m mirrors, shaped to provide one continuous primary mirror<br />

surface. Its collecting area would be equivalent to a 21-m single round primary.<br />

All <strong>of</strong> <strong>the</strong>se projects face a number <strong>of</strong> technical hurdles to work well enough to justify<br />

<strong>the</strong>ir cost. We have already mentioned that <strong>the</strong>ir sensitivity gains are dependent on <strong>the</strong><br />

success <strong>of</strong> Multi-Conjugate Adaptive Optics. Their large downward looking secondary<br />

mirrors are a challenge to mount, because <strong>the</strong>y have to be “hung” against <strong>the</strong> pull <strong>of</strong><br />

gravity, a much more difficult arrangement than is needed for <strong>the</strong> upward looking<br />

primary mirrors. For <strong>the</strong> segmented designs, <strong>the</strong> electronic control loop to maintain<br />

alignment will be very complex. All <strong>of</strong> <strong>the</strong>m will be severely challenged by wind, which<br />

can exert huge forces on <strong>the</strong>ir immense primary mirrors and structures. None<strong>the</strong>less, we<br />

can hope that <strong>the</strong> financial and technical problems will be surmounted and that <strong>the</strong>y will<br />

become a reality.<br />

References<br />

Epps, H. W., & Fabricant, D. 1997, AJ, 113, 439<br />

Fabricant, D. et al. 2004, SPIE, 5492, 767<br />

Wynne, C. G. 1974, MNRAS, 167, 189<br />

Fur<strong>the</strong>r Reading<br />

Bely, The Design and Construction <strong>of</strong> Large Optical Telescopes, 2003. Covers<br />

management principles as well as <strong>the</strong> usual topics in instrumentation.<br />

Bracewell, R. N. 2000, The Fourier Transform and its Applications, 3 rd Ed.<br />

Press, W. H. et al. 2007, Numerical Recipes, 3 rd Ed.<br />

Schroeder, Astronomical Optics, 2 nd edition, 1999. Ra<strong>the</strong>r ma<strong>the</strong>matical, but classic<br />

treatment <strong>of</strong> this topic.<br />

Wilson, Reflecting Telescope Optics, 2 nd ed., 1996; probably more than you wanted to<br />

know

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!