Sie sind auf Seite 1von 57

11/5/12

Image Quality Metrics

The US government requires a license to distribute imagery taken by a commercial


satellite. The license normally specifies the allowable resolution, defined as size of the pixel
projection on the ground. For the WorldView 1 and 2, and the GeoEye 1 satellites the license
allows for the distribution of half-meter resolution imagery, even though the satellite was capable
of marginally better resolution. Using resolution as the sole criterion for a license is somewhat
surprising because image quality also has a strong dependence on the aperture configuration, and
the signal to noise.

As early as the 1950s photo interpreters used an empirical equation to compare image
quality between sensor designs. That equation included not only resolution but also a metric
related to the point spread function. The 1980s saw the development of a more sophisticated
approach in which image quality was expressed in NIIRS, which image analysts compare to a
rating scale concerning interpretability. The General Image Quality Equation (GIQE) has been
developed, which makes it possible to compute the NIIRS as a function of three main factors: (1)
resolution (expressed in terms of the ground sample distance (GSD), (2) the relative edge
response (RER), which is indirectly related to the point spread function, and (3) the signal to
noise ratio (SNR). The form of the GIQE is

NIIRS= c0 + c1 ⋅ log10 GSD + c2 ⋅ log10 RER + c3 ⋅ H + c4 ⋅ G/SNR

H is an overshoot value derived from the edge response function, and G is a noise gain
associated with image post processing. The c i are constants. So for example, a change of ± one

NIIRS corresponds to halving or doubling the distance between the sensor and the scene.

Determination of the RER requires a frequency-based approach, using modulation


transfer functions (MTFs) to model € the effects of sampling, diffraction, smear, jitter, and
wavefront aberrations. Specific results are presented, without going into their formulation using
Fourier transforms. Comparisons are provided for different combinations of sampling and
diffraction for three “ideal” systems. Simulated imagery examples are provided in some
instances.

The design of a satellite imaging system to achieve a specified NIIRS involves choosing
sensor parameters for which the combination of GSD, RER and SNR, when inserted into the
GIQE, provide the required NIIRS. That process is described herein. Telescope and detector
design are not covered, though this process can provide requirements for these elements. The
analysis pertains only to incoherent imaging, which is what all imaging satellites employ. The
intent is to provide analytic tools by which to compare designs at the conceptual level.

1 R. R. Auelmann
11/5/12

Optical Imaging

Optical imaging of the Earth from a satellite is primarily using either reflected sunlight
off the scene or the thermal emission of the scene. With the advent of lasers one might also
envision the imaging of reflected laser radiation of the scene. The optical band is normally
defined as radiation wavelengths between 0.3 µm and 15 µm. This spectrum can be divided into
the following subdivisions:

0.2 µm to 0.4 µm: photographic ultraviolet (UV)


0.4 µm to 0.7 µm: visible (V)
0.7 µm to 1 µm: near infrared (NIR)
1 µm to 2.7 µm: short wave infrared (SWIR)
________________________________________
2.7 µm to 6.3 µm: mid wave infrared (MWIR)
6.3 µm to 15 µm: long wave infrared (LWIR)

(Note there is not precise agreement among references for the transition wavelengths between
spectrum subdivisions.) For the first four bands that use reflected sunlight, the solar irradiance
and the scene reflectance are critical. For the latter two bands that use the thermal emission, the
scene temperature and emissivity are critical.

Figure 1 shows the solar spectral irradiance both at the top of the atmosphere and at sea
level when the Sun is directly overhead. Note that when the Sun is represented as a 5250°K
blackbody the corresponding irradiance approximates the actual irradiance at the top of the
atmosphere. The various constituents in the atmosphere cause transmission losses, so only part of
the solar irradiance reaches the surface of the Earth (the dark portion part of the spectrum).

2500

2000
Spectral Irradiance (W/m2/µm)

1500

1000

500

0
0.5 1.0 1.5 2.0 2.5
Wavelength (µm)

Figure 1. Solar Spectral Irradiance (adapted from Infrared Handbook)

2 R. R. Auelmann
11/5/12

Most photography is done in the visible spectrum (0.4 µm – 0.7 µm) within which the
human eye is sensitive. The Bayer color filter array shown in Figure 2 is a popular detector
format. It employs half the number of pixels in green, which is the portion of the spectrum most
sensitive to the human eye. The red and blue pixels each count for one quarter of the total.

Figure 2. Bayers Filter

The Bayers filter is not used for satellites imagers. Rather most satellite sensors operate
in visible and near infrared (VNIR) portion of the spectrum roughly 0.4 µm – 0.9 µm, and
employ a high-resolution panchromatic (pan) band with imagery portrayed in grey levels. For
some satellites (WorldView 1, for example) this is the only band. But more often there are lower
resolution multi-spectral (MS) bands, three of which are blue-green-red and sometimes a fourth
near infrared (Figure 3). Later we describe how high resolution natural-color images can be
produced using a combination of the high-resolution pan image and low-resolution red-green-
blue images. The technique is called pan sharpening. The lower resolution multi-spectral bands
are also useful for other purposes such as determining the health of crops and vegetation. For the
most part, we will be concerned with the image quality of the high-resolution pan band.

Figure 3. High Resolution Pan Band with Lower Resolution Multi-Spectral Bands

The useful portions of the spectral band depend on three factors: atmospheric
transmission, sensitivity of the detector measured in terms of the quantum efficiency (the
efficiency of converting photons into electrons), and the solar spectral irradiance profile (see
Figure 1). Atmospheric transmission values for different portions of the optical spectrum are
shown in Figure 4. The relevant (orange) profile is the vertical transmission through the entire
atmosphere. The other (black) profile is for horizontal transmission at sea level.

3 R. R. Auelmann
11/5/12

Figure 4. Atmospheric Transmission

Except where specifically noted (as for existing satellites) performance evaluations will
be made in four spectral bands:
VNIR (0.4 µm – 0.9 µm)
SWIR (0.9 µm – 1.7 µm)
MWIR (3 µm – 5 µm)
LWIR (8 µm – 12 µm)

The VNIR/SWIR cutoffs do not precisely match the previous transition points is due to the
quantum efficiency spectral profiles of the available detector materials. So for example, silicon
detector response for the VNIR band typically cut off near 0.9 µm, while the SWIR band is
largely defined by the response of Indium Gallium Arsenide (InGaAs) detectors.

4 R. R. Auelmann
11/5/12

Sampling.

Initially, satellite images of the Earth and the clouds were collected either on film or
using television cameras, but since the late 1980s most satellite imagery has been collected in
digital format using solid state detectors. A digitized image can be thought of as a two-
dimensional matrix of picture elements, called “pixels”. Each pixel is located by its row and
column indices, which generally correspond to elevation and azimuth, respectively. These pixels
become apparent when the image is highly magnified as shown in Figure 5.

Jefferson Memorial -IKONOS II 10/18/99 (Space Imaging)

Figure 5. Magnification Shows Individual Pixels (portion of Jefferson Memorial Image above)

5 R. R. Auelmann
11/5/12

Assume for the moment that we are dealing with a frame camera in which a two-
dimensional array of detectors is located in the focal plane. Such an array is depicted in Figure 6
where the individual detector dimensions are dx and dy, and the center to center spacing between
detectors are px and py, which define the sampling intervals. Most detector arrays are so tightly
packed that the detector dimension is equal to the sampling interval, in which case the fill factor
is 100% as is the figure on the right. Henceforth, unless otherwise noted we shall assume that the
pixel spacing is equal to the detector dimension.

dx px

dy
py

Figure 6. Detector Array Geometry.

The detector elements are sampled (readout) at periodic intervals ts. Their outputs in
photoelectrons correspond to the brightness output of each detector element. The value assigned
to each pixel corresponds to the brightness of the image at a specific pixel location. These values
can be stored in a computer as binary numbers. The number of bits assigned to each pixel
determines the number of gray scale levels L, which is equal to 2 raised to the b power where b
is the number of binary bits. With one bit per pixel, the pixel can only be black for 0 or white for
1. With two bits per pixel we can have four gray scale levels and for three bits per pixel we can
have eight gray scale levels as depicted in Figure 7.

000 001 010 011 100 101 110 111

Figure 7. Eight Gray Scales With 3-bit Radiometric Resolution

Images in which the intensities over a broad spectral band of colors are represented as
gray tones ranging between black and white are called “panchromatic” images. Figure 8
compares between 2-bit and 8-bit panchromatic imagery. The panchromatic imagery provided by
the SPOT satellites has 6-bits per pixel radiometric resolution, providing 64 gray scale levels,
which sounds like a lot. Actually it is not. The dynamic range in brightness of real scenes can be
huge, making it difficult, for example, to detect objects in shadows with only 6-bit resolution.
That is why panchromatic imagery provided by the newer commercial imaging satellites such as
IKONOS II provide 11 bits per pixel with a corresponding 2048 gray scale levels.

6 R. R. Auelmann
11/5/12

Figure 8. Comparing 2-bit and 8-bit Images.

Also high radiometric resolution requires either detectors with large well capacity to store
photoelectrons or special high contrast film. The easiest way to increase well capacity is to
increase the area of the detector, which means increasing the pixel size. The size of the pixel and
the focal length of the camera, define the highest spatial frequency portrayed in the image.

Resolution.

Consider a simple pinhole camera (Figure 9) consisting of an enclosed box with a small
hole in one end, a detector array on the opposite wall, and a shutter located at the pinhole.

Δ
r

Figure 9. Pinhole Camera.

7 R. R. Auelmann
11/5/12

When the shutter is opened, an inverted image of the scene is projected on to the back
wall where the detector array is located. After a specified integration time ts, the shutter is closed
and an image is captured. The scene resolution Δ, corresponding to the pixel spacing p is given
by
Δ = r ⋅ (p/ f )

The detector instantaneous field of view (IFOV) is equal to d/f. With few exceptions d and p are
equal, in which case
€ Δ = r ⋅ IFOV

Ground Sample Distance (GSD)



A more widely used parameter for representing resolution is the ground sample distance
(GSD), which takes into account that the ground is not always viewed from directly overhead.
Figure 10 depicts a ground plane (yellow) centered at the LOS aim point, and intersecting plane
(blue) plane normal to the LOS (the z-axis). The zenith angle ψ is between the local vertical
(LV) and z. The axes x and y are projections of the detector sampling axes in the blue plane. The
angle between the intersection of the two planes and the x-axis is γ .

Figure 10. Intersection of Ground Plane and Pixel Projection Plane at the Aim Point

The projections of the pixel sampling distances along the x and y axes are

px py
Δx = r ⋅ , Δy = r ⋅
f f

The projections of these into the ground plane define the x and y components of the GSD:

px sin 2 γ py cos 2 γ
GSDx = r ⋅ cos2 γ + , GSDy = r ⋅ sin 2 γ +
f cos 2 ψ f cos 2 ψ

€ 8 R. R. Auelmann
11/5/12

and the GSD is the geometric mean of these components: GSD = GSDx ⋅ GSDy . In most
instances px = py = p so that the GSD becomes

14
p +% 2 €2 γ ( % 2
sin cos2 γ (.
GSD = r ⋅ -' cos γ + * ⋅ ' sin γ + *0
€ f ,& cos 2 ψ ) & cos2 ψ )/

For the often encountered conditions when γ = 0 or 90° the GSD reduces to
€ r ⋅ (p/ f )
GSD =
€ cosψ

Recall that when the pixel sampling distance is equal to the pixel dimension p / f is the IFOV .

This last form is an excellent approximation for most viewing geometries as can be seen
from Figure 11, which compares the exact and approximate expressions for GSD, normalized on
r ⋅ ( p / f ) . The differences are largest when γ = 45° . The maximum€difference is 3% at ψ = 45°
and grows to 11% at ψ = 60°.

€ € €

Figure 11. Differences Between Exact and Approximate Forms of GSD

The zenith angle dependence favors viewing from overhead, rather than from the side.
However, there are clearly instances when viewing more from the side preferable to imaging
from directly overhead. This is the case when attempting to classify and identify ships, and there
are other instances as well.

9 R. R. Auelmann
11/5/12

Optics.

The required integration (exposure) time depends on the brightness of the scene, the
sensitivity of the detector, and the size of aperture. Indeed, integration time is inversely
proportional to the aperture area. Because image smear due to camera pointing errors varies
directly with integration time, long integration times are undesirable. If one enlarges the pinhole,
so as to reduce the integration time, they encounter another problem. The larger pinhole allows
rays from a point of the scene to reach different points in the image plane. When this spread of
the incoming rays approaches the pixel size, the image becomes blurred. This problem can be
solved by the addition of a lens, which focuses rays emanating from a point on the object to a
corresponding point of the image plane. The lens allows an increase in aperture size and a
corresponding reduction in the required integration time. But a single lens is sufficient only for
the on-axis rays. As the field of view (FOV) increases one requires lens combinations (e.g. the
Petzval lens configuration on the CORONA satellites).

There are two practical limitations to the use of lens. First, being refractive elements they
bend rays different amounts depending on the wavelength. And second, the weight of the lens
becomes excessive as the size grows. That is almost all high-resolution imaging satellites use
reflective elements (mirrors) rather than lens. The most commonly used telescope configuration
used by modern commercial imaging satellites is the Three-Mirror Anastigmat (TMA) like the
one illustrated in our discussion of the Pleiades satellite.

Integration time is not the only factor driving aperture size. Diffraction, an effect due to
the wave nature of light that causes the light to spread when it passes through an aperture, can
dictate the minimum size for the aperture if the scene is at long range. Consider a point source of
radiation at a distance r from a circular aperture. Assume that r is so much larger than the
aperture diameter D that all the rays passing through the aperture are nearly parallel, certainly a
valid assumption for satellite imaging. Then, the intensity pattern in the image plane is circular
with a bright central lobe and fainter side lobes (Figure 12).

Figure 12. Diffraction pattern for a distant point source imaged through a circular aperture
(www.matter.org.uk/tem/resolution.htm)

The intensity pattern is known as the point spread function (PSF). Figure 13 shows the
cross section of this particular PSF, normalized on λ/D. Mathematically it is described by a

10 R. R. Auelmann
11/5/12

Bessel function of the first kind. The first null is at radius 1.22 λ/D where D is the aperture
diameter. The second null is at radius 2(1.22 λ/D), the third null is at radius 3(1.22 λ/D), and so
on. Nearly 84% of the total energy is contained within the central lobe. The full width at the half
maximum (FWHM) of the PSF for an ideal circular aperture is equal to 1.028 λ/D (often
approximated as λ/D). Its relationship to IFOV is of particular interest.

Figure 13. Point Spread Function (PSF) for a Circular Aperture

Optical Q.

Optical Q defines the balance between sampling resolution and diffraction. Let λmean
denote the mean wavelength of the spectral band. Then in linear units Q is the ratio of F ⋅ λmean
to the size of the pixel p. In angular terms it is the ratio of the optics resolution λmean /D (set by
diffraction) and the sampling resolution defined by the IFOV:

fλ /D F ⋅ λ λ /D €
Q= = = €
p p IFOV

A sensor with Q less than one has its resolution limited by spatial sampling, while a sensor with
Q greater than one has its resolution limited by diffraction. The two components of Q are
depicted in Figure 14. The € latest US commercial imaging satellites have Q near 1. The important
a roll is played by this single parameter Q cannot be over emphasized.

11 R. R. Auelmann
11/5/12

Figure 14. Components of Optical Q

Effective Resolution

Recall that licenses for US commercial satellites have limits on the nadir GSD (simply
the product of the IFOV and altitude). But GSD alone is insufficient to represent the apparent
resolution of even an “ideal” system, one that is free of optical aberrations, smear and jitter. The
effective resolution should also include the diffraction impact of “perfect” optics (represented by
the upper portion of Figure 14). Camera manufacturers have long realized this even back in the
film era when they devised means of measuring film resolution and lens quality in units of line
pairs per millimeter (lp/mm), and proposed models for combining the two quality measures into
an effective sensor resolution (also measured in lp/mm). Perhaps the earliest model was

1 1 1
= +
Rsensor Roptics R film

where Rsensor is the sensor resolution. The attraction of this approach is that the optics and film
resolution could be separately measured using bar charts and combined using the model to
€ For digital cameras with diffraction limited optics one simply
determine the sensor resolution.
sets
€ R film = Rdetector = 1/(2 p)

D 1
Roptics = =
2 ⋅ f ⋅ λmean 2 ⋅ F #⋅ λmean

where F # = f /D . The factor of two is because the measurements are in line pairs rather than
lines. It follows that

12 R. R. Auelmann

11/5/12

1 $λ '
= 2 ⋅ ( F #⋅ λmean + p) = 2 f ⋅ & mean + IFOV )
Rsensor % D (

and the corresponding effective angular resolution is


€ $λ ' λ $ 1'
α eff = & mean + IFOV ) = mean ⋅ &1+ )
% D ( D % Q(

where Q is the ratio of λmean /D to IFOV. This equation is the one used by photo interpreter’s
equation to model the proposed the Corona satellite (circa 1958). And it is currently used in Fuji
camera data sheets. €
So we will refer to this as the Fuji model.

Unfortunately, it is an empirical equation that is not universally accepted. Kodak camera
data sheets are based on a different model:

1 1 1
2
= 2
+ (Kodak)
R sensor R optics R 2film

which yields the following equation for effective angular resolution:


€ 1/ 2
$ λ '2 λ $ 1'
α eff = & mean ) + IFOV 2 = mean ⋅ &1+ 2 ) (Kodak)
% D ( D % Q (

Obviously both Fuji and Kodak models cannot both be correct and indeed both are off the mark.
€Using the GIQE it can be shown that the desired solution for effective angular resolution
is given by
IFOV λmean /D
α eff = =
RER Q ⋅ RER

where RER is the yet to be defined Relative Edge Response. The empirical equation (analogous
to the Fuji and Kodak models, but with a different exponent) that provides an almost perfect
match to the GIQE derived€value for α eff is given by

1 1 1
1.35
= 1.35
+ 1.35
R sensor Roptics R
detector

which yields
1/1.35
λ % 1 (
α eff = mean ⋅ '1+ 1.35 *
€ D & Q )

Numerical values for Q ranging between 0.5 and 2 are presented in Table 1 and graphed in
Figure 15. The results are normalized on λmean /D rather than on IFOV, because most satellite

13 R. R. Auelmann

11/5/12

sensors are limited by the aperture size. It is clear the empirical model with an exponent of 1.35
is much closer match to the GIQE based “truth” model than either the Fuji or Kodak models.

TABLE 1. GIQE Model and Empirical Models

αeff/(λ/D)
Q RER GIQE Fuji Kodak Proposed
0.5 0.747 2.677 3 1.732 2.556
1 0.600 1.667 2 1.414 1.671
1.5 0.475 1.404 1.667 1.291 1.402
2 0.383 1.305 1.5 1.225 1.278

Figure 2. Comparing Empirical Models with GIQE Derived Model

14 R. R. Auelmann
11/5/12

Images of Washington DC one taken on the first Corona KH-4B flight and the other by
IKONOS II were compared in Section 1. Table 2 lists the effective ground resolution. It should
be noted that these values represent idealized conditions. The model does not account for the
significant smear likely present in the KH4B image.

TABLE 2. “Effective” Ground Resolution for Two Satellites

Corona KH4B Ikonos II


λmean/D, (mr) 0.0034 0.0010
IFOV, (mr) 0.0051 0.0012
Optical Q 0.67 0.80
F 3.5 14.3
Focal length, f (m) 0.61 10
Altitude, r (km) 150 682
Roptics (lp/mm) 280 52
Rfilm or Rdetector (lp/mm) 160 42
Rsensor (lp/mm) 120 27
Resolution, Δ = r•IFOV (m) 0.77 0.82
Eff. angular resolution, αeff (mr) 0.0072 0.0018
Eff. ground resolution, ΔT (m) 1.07 1.24

The “effective” sensor angle and nadir ground resolutions for the pan band of all the
previously considered French and US commercial imaging satellites are presented in Table 3.
The pixel pitch values cited for SPOT 5 and Pleiades are based on Supermode processing. The
actual pixel pitch values are 6.5 µm and 13 µm, respectively. Pleiades and Ikonos 2 are at nearly
the same altitudes, yet Pleiades achieves 20% better effective ground resolution despite a smaller
aperture. Based on the standard definition WorldView 1 has an18% lower ground resolution than
its predecessor QuickBird 2, which almost identical telescopes. Yet based on “effective” ground
resolution the advantage is only 6.4%, again illustrating why the standard definition for ground
resolution is inadequate. Also note that the GeoEye 1 and WorldView 2 satellites have identical
pan sensors, but are at different orbit altitudes.

TABLE 3. Effective Sensor Angle and Ground Resolutions (8 satellites)

French Satellites US Satellites


SPOT 1- 4 SPOT 5 Pleiades Ikonos II QuickBird 2 WorldView 1 GeoEye 1 WorldView 2
Mean Wavelength, λmean (µm) 0.615 0.615 0.655 0.675 0.675 0.675 0.675 0.675
Aperture diamater, D (m) 0.31 0.31 0.65 0.70 0.6 0.6 1.1 1.1
λmean/D, (µr) 1.98 1.98 1.01 0.96 1.13 1.13 0.61 0.61
Pixel pitch, p (µm) 13 4.6 9.12 12 12 8 8 8
Focal length, f (m) 1.08 1.08 12.905 10 8.84 8.00 13.3 13.3
IFOV, (µr) 12.0 4.3 0.7 1.2 1.4 1.0 0.6 0.6
F# 3.48 3.48 19.85 14.29 14.73 13.33 12.09 12.09
Optical Q 0.165 0.466 1.426 0.804 0.829 1.13 1.02 1.02
Eff. angular res., αeff (mr) 12.8 5.3 1.4 1.8 2.1 1.8 1.0 1.0
Altitude, h (km) 828 828 693 682 451 497 685 773
Resolution, Δ = r•IFOV (m) 10.0 3.5 0.49 0.82 0.61 0.50 0.41 0.46
Eff. ground res., ΔT (m) 10.6 4.4 1.00 1.24 0.94 0.88 0.70 0.78

15 R. R. Auelmann
11/5/12

Before leaving the subject of resolution, it is well to point out that under special
circumstances it is possible to detect elements (lines in particular) that are much smaller in width
than the ground resolution. Figure 16 is a natural color image of tennis courts collected by the
QuickBird 2 satellite with a GSD resolution of 61 centimeters and an effective resolution of 88
centimeters. Line widths on a tennis court are 5 centimeters, an order of magnitude smaller than
the resolution, yet they are clearly visible. This is mainly due to the large difference in brightness
between the lines and the background.

Figure 16. Tennis Courts Collected by QuickBird 2 at 0.61-m Resolution.(DigitalGlobe).

16 R. R. Auelmann
11/5/12

Frequency Domain Analysis

Frequency is normally understood to mean the number of cycles per unit time. In other
words, it is a temporal measure. Optical design also uses frequency, but in a spatial rather than
temporal sense. Because image space is two-dimensional, we speak of two-dimensional spatial
frequencies yet often the spatial frequency content is the same in both directions. If a scene has
many sharp edges, the image is said to rich in high spatial frequency content. If there are no hard
edges, and only soft transitions, the spatial frequency content is low.

Consider the one-dimensional scene (blue) in Figure 17. It has a single frequency. The
vertical axis is the scene intensity J and the horizontal axis is the direction in which the scene
intensity is modulated. The modulation of a scene is the relative deviation of the intensities J
from their average value as given by

J dev J max − J min


Mod(O) = =
J ave J max + J min

How well can an imaging system reproduce the scene? Provided the sensor has sufficient
resolution, it will be able to measure the scene frequency. However, the modulation amplitude of
the image
I − Imin
Mod(I) = max
Imax + Imin

will almost always be of lower amplitude than the scene.


scene image
J max
I max
Intensity

Jave

I min
J min

Figure 17. Scene and Image Modulations.

Using Fourier transforms, both the scene and the image can be represented as a
summation of sine wave patterns for various spatial frequencies. We refer to these as the scene
and image spectra modulation. The ratio of the image modulation to the object modulation is
called the modulation transfer function (MTF). It represents how well the imaging system is able
to reproduce the modulation present in a scene. And it depends on all the operations (sampling,
diffraction, smear, jitter, aberrations, etc.) involved in forming the image. Each of these
operations can be represented by its MTF. The advantages of this frequency-based approach are
three-fold. First, the combined or system MTF is simply the product of the MTFs of the

17 R. R. Auelmann
11/5/12

constituent elements. Second, the constituent MTFs can be analytically determined, which aids
in the design of the imaging system. Third, both the constituent and system MTFs can be
measured after the elements are built.

The two most important contributors to the system MTF are (1) the sampling of the
detector/telescope combination, and (2) the diffraction characteristics of the telescope.
Depending on the optical Q, either sampling or diffraction determines the cutoff frequency of the
system (the lowest spatial frequency at which the MTF goes to zero).

Sampling MTF. The sampling cutoff frequency depends on its pixel spacing p and the
telescope focal length f as follows
f 1
χ sampling = =
p IFOV

The sampling MTF is given by


sin(πχ )
€ MTFsampling =
πχ

where χ is the frequency normalized on χ sampling . This function is plotted in Figure 18.

Figure 18. Sampling MTF (normalized on the cutoff frequency 1/IFOV).

Aliasing. When the scene contains spatial frequencies higher than half the sampling
frequency, artifacts, called aliasing errors, appear in the image, which are not present in the
scene. This phenomenon is peculiar to periodic sample data systems, and occurs because the
scene frequencies beat with the sample frequency. The insidious nature of aliasing is that,
without a ‘truth” image to compare, there is no way to tell if aliasing is present in an image.

18 R. R. Auelmann
11/5/12

In order to avoid aliasing either the system modulation transfer must be zero at
frequencies above the Nyquist frequency (defined as half the sampling frequency) or the scene
must contain no frequencies above the Nyquist frequency (see Figure 10). This is the so-called
Nyquist sampling criterion. Figure 19 illustrates the situation when the scene frequency is
exactly equal to the Nyquist frequency; that is, there are two samples per scene cycle.

Figure 19. Sampling at the Nyquist Frequency. Sample interval is half the scene cycle interval.

What happens when this criterion is not satisfied? Consider the situation in Figure 20.
Here there are only six samples within the four-cycle scene, which means that the scene
frequency is 3/2 times the Nyquist frequency. The aliased image is the lowest frequency pattern
consistent with the sampling. In this case it is at half the frequency of the real scene. The aliased
image appears at a frequency as far below the Nyquist frequency as the scene frequency is above
the Nyquist frequency.

Figure 20. Example of Aliasing.

19 R. R. Auelmann
11/5/12

It is as though the portion of the MTF above the Nyquist frequency flips these high
frequency scene inputs into lower frequency image outputs as depicted in Figure 21 showing the
total system MTF with a cutoff past the Nyquist frequency.

System MTF
Nyquist

Sampling

Frequency

Figure 21. Aliasing Folds the MTF Back About the Nyquist Frequency.

The early imaging satellites (e.g. MSS, Landsat) employed discrete scanning detectors
with electronic anti-aliasing filters designed to reduce the MTF response at frequencies above the
Nyquist frequency. Such filters acted upon the detector signals prior to sampling. The anti-
aliasing filter used on the Landsat Thematic mapper sensor reduced the in scan and cross scan
MTFs at high frequency but did not completely eliminate aliasing.

Most commercial imaging systems now use CCD arrays in which the sampling is
performed on chip in the spatial domain, in which case electronic anti-aliasing filters cannot be
used, because the filtering must occur prior to sampling. One way to filter prior to sampling is to
design the optics so that the diffraction cutoff frequency is at the Nyquist frequency. However,
none of the current commercial imaging systems avail themselves of this approach. To the
degree that the system MTF has gain beyond the Nyquist frequency, they will exhibit aliasing.

Oddly enough, film systems are immune from aliasing. Photographic film contains tiny
crystals of silver halide salts ranging between 0.2 and 2 µm in size. The slower the film, the
smaller are the size of the crystals. These particles are the fundamental light sensitive elements.
When the film is developed they gather into randomly distributed clumps of metallic silver called
“grains”. It is these grains (represented in Figure 22) that are measured when determining film
resolution. The average spacing p between grain centers, along orthogonal directions, is equal to
the reciprocal of twice the film resolution (e.g. for a film resolution of 160 lp/mm, p is 3.125
µm). This spacing is equivalent to pixel pitch in a digital system. Because of the random grain
distribution there is no “beating” with the scene frequencies as in periodically sampled CCD
systems.

Figure 22. Representation of the random distribution of film grain

20 R. R. Auelmann
11/5/12

MTF for Circular Aperture. The diffraction cutoff frequency depends only on the
aperture shape and the mean wavelength λ mean . For a circular aperture telescope of diameter D
the cutoff frequency is given by
D
χ diffraction =
λmean

For a filled circular aperture (no obscuration) the telescope MTF is given by
€ 2
(
MTFdiffraction = cos−1 ρ − ρ 1− ρ 2
π )
where ρ is the relative frequency normalized on the diffraction cutoff frequency. The MTF
function is plotted in Figure 23.

Figure 23. Diffraction MTF for a Filled Circular Aperture. Note that in this plot χ is normalized
on the diffraction cutoff frequency rather than the sampling frequency.

There is a useful geometric interpretation to this curve. The values along the curve
correspond to the relative area overlap of two unit-diameter circles depicted at the top of the
figure. When the displacement between the circles is zero, the overlap is equal to the area of the
circle, which when normalized to one corresponds to the MTF at zero frequency. When the
displacement between the two circles is equal to half their diameter, the overlap area is 0.391,
which corresponds to the MTF at a frequency of a half. Finally when the displacement is equal to

21 R. R. Auelmann
11/5/12

the diameter of the circle there is no overlap, which corresponds to the zero value of the MTF at
a frequency of one. This interpretation can be used to evaluate the MTF for virtually any aperture
shape.

Many imaging systems employ circular apertures with a central obscuration. This is
generally the case for reflective telescopes that are axial symmetric. Let ε denote the ratio of the
linear obscuration to the outside aperture diameter. Then the diffraction MTF expressed in terms
of the frequency ρ normalized on the diffraction cutoff frequency D/ λ is given by the following
set of equations.

€ €

The diffraction MTF is plotted in Figure 24 for ε = 0, 0.2, 0.5 and 0.8. The effect of
increasing ε is to increase the higher spatial frequencies at the expense of the low and mid spatial
frequencies. As such it represents a form of edge sharpening. For ε ≤ 0.2, which is representative
of most telescopes, the effect of the central obscuration on MTF is minimal. For ε ≥ 0.8 we are

really dealing with an annular aperture and effect of the obscuration is extreme.


22 R. R. Auelmann
11/5/12

1.0

0.9

0.8

0.7

0.6
MTF

0.5

0.4

0.3

0.2

0.1

0.0
0 0.2 0.4 0.6 0.8 1
Relative Frequency, ρ

Figure 24. MTFs for Circular Apertures with Central Obscurrations

Hexagonal Mirror Segments. For large primary mirrors it may be more effective to employ
hexagonal mirror segments rather than a single large circular mirror. The dimensions of a hex segment
are shown in Figure 25, where s is the side length of a segment. The area of a segment is

3 3 2
Aseg = s = 2.5981⋅ s2
2

Figure 25. Hex Segment Dimensions

Figure 26 shows N = 6, 18 and 36 segment configurations. The John Web Space Telescope being
developed by NASA has 18 segments. The fill factor is compared to both the area of the circumscribed
circle of diameter Dmax and to the mean diameter Dmean and increases as N increases. The mean diameter
Dmean is the geometric mean Dmax Dmin . Observe that the ratios Dmin /Dmax and Dmean /Dmax are
independent of N. If a segment were placed in the central obscuration the only thing that would change
is the fill€factor which would increase by€(N + 1)/N.
€ € € €

23 R. R. Auelmann
11/5/12

Figure 26. Three Hex Mirror Segment Options

The dimensions Dmax and Dmin refer to the max and min spatial cutoff frequencies of the aperture MTF.
Figure 27 shows the orientations of the max and min diameters for N = 36. The green axes denote the
orientations of the max cutoff frequencies and the red values denote the orientations of the min cutoff
frequencies. For each max orientation there is a min orientation 90° removed. For evaluating image
quality€it is useful€to employ an “equivalent” circular aperture of diameter Dmean and central obscuration
with fill factor η fill(mean ) (Figure 27) and compute diffraction using the model of Figure 24.


Figure 27. Orientation of the Max and Min Diameters (N = 36) and Equivalent Circular Aperture

24 R. R. Auelmann
11/5/12

MTFs for Smear and Jitter. If the sensor LOS moves across the scene during the
integration time, the resulting image will be blurred. It is necessary to distinguish between two
types of motion, because they have different effects on image degradation. If the camera LOS
translates at a nearly constant rate across the scene during the integration time, the motion is
termed smear. If, on the other hand, it oscillates during the integration time, the motion is termed
jitter. Both of these effects can be expressed in terms of their MTFs. The MTF for smear is

sin(π k sχ ) θ˙LOS t int


MTFsmear = , ks =
π ksχ IFOV

where χ is normalized on the sampling frequency. The constant ks is the change in LOS
direction during the integration interval t int normalized on the IFOV. Thus ks = 1 corresponds to
one pixel of smear. €
The smear MTF is a sinc function, which for ks = 1 is identical to the
sampling MTF. Moreover, for scanners, ks = 1 corresponds to the smear in the along-scan
€ €
direction produced by the scan process itself.
€ €
The MTF for jitter is €

σj
[
MTF jitter = exp −2(π k j χ ) ,
2
] kj =
IFOV

Here σ j is the rms jitter amplitude. The MTFs for smear and jitter are depicted in Figure 28.

1.0 €
1.0

k s = 0.2 k j = 0.1

0.8 0.8

k s= 0.5
0.6
Smear MTF

0.6
Jitter MTF

k j = 0.2
ks = 1
0.4 0.4
k j = 0.5
0.2 0.2

kj= 1
0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Relative Frequency Relative Frequency

Figure 28. Smear and Jitter MTFs. The relative frequency in these plots is normalized on the
sampling frequency 1/IFOV.

25 R. R. Auelmann
11/5/12

The MTF is two to three times more sensitive to jitter than to smear (i.e. comparing the
sensitivities to ks and k j ). This is fortunate because for short exposures (say a few
milliseconds), most disturbances appear as smear rather than jitter.

MTF for Wavefront Aberrations. Ideally the sensor optics are designed such that a

point source€in the in the object plane is projected as a spherical wave to the focal plane.
Deviations in the wavefront are referred to as wavefront aberrations. The primary sources of
wavefront aberration include: (1) inherent limitations in the optical configuration; (2) mirror
fabrication errors; (3) focus errors; and (4) misalignment of the optical elements due to dynamic
disturbances. Often these aberrations are modeled in terms of Zernike polynomials, in which
case they are classified terms of defocus, astigmatism, coma, spherical, and higher order terms.
Almost all imaging systems employ some means of active focus control. And large telescopes
sometimes include active wavefront correction where the aberrations are sensed and
compensated using an actively “deformable” mirror in what is termed “adaptive optics”.

Obviously, the process of accurately modeling the MTF losses associated with wavefront
aberrations is dependent on a large number of design inputs. Indeed, in some cases the MTF is
not even the appropriate representation. This is because some aberrations yield complex values
(in the sense of √-1), requiring the use of the optical transfer function (OTF) rather than its
absolute value the MTF. This subtlety is ignored in the general definition of image quality.

In the early stages of system modeling it is not uncommon to include a “place holder”
MTF to represent the effects of wavefront aberrations. One example is

-/ & 2 1/
2πδ )
MTFwavefront = exp.−( [
+
/0 ' λmean *
1− exp( − ρ 2
/]ρ ripple ) 2
2

/3
δ = rms ripple amplitude
ρ = frequency normalized on the diffraction cutoff frequency
ρ ripple = ripple correlation scale

The normalization is logically with respect to the diffraction cutoff frequency because the errors
are associated with the optics.

For a diffraction-limited telescope the PSF (see Figure 13) has a peak value of one. At the
radial distance ρ = 0 the MTF reduces to the Strehl ratio:

- ' *20
2πδ
SR = MTF ( ρ = 0) = exp/−) ,2
€ /. ( λmean + 21

The SR defines the peak value of the PSF in the presence of wavefront aberrations. Normally the
optics are designed so that the SR ≥ 0.8 which results when the ripple amplitude corresponds to
€ The frequency dependence enters the MTF through the factor
0.075 waves of distortion.

26 R. R. Auelmann
11/5/12

[1− exp(−ρ 2 2
]
/ ρ ripple ) . Figure 29 shows the MTF that corresponds to an SR = 0.8 and which has
the ripple correlation scale equal to 0.1 times the diffraction cutoff frequency. Note that for
frequencies higher than 20% of the diffraction cutoff, the value of the MTF is equal to the SR.
Finally note that this MTF must be rescaled to the sampling cutoff when it is included in
€ computing the system MTF.

Figure 29. Generic Wavefront MTF

27 R. R. Auelmann
11/5/12

Three Ideal Systems. As previously noted, the system MTF is obtained by taking the
product of the constituent MTFs. But in order to do this the MTF components must be
normalized to a common cutoff frequency. As noted above, the cutoff frequencies for sampling
and diffraction are not the same. And as shown later, the cutoff frequencies for smear and jitter
are logically normalized on the sampling cutoff frequency, while the cutoff frequencies for
mirror aberrations are logically normalized on the diffraction cutoff frequency. In computing the
system MTF we choose to normalize on the sampling MTF. The choice is based on how the yet
to be defined edge response function is computed.

Consider now an idealized system with only sampling and diffraction for a filled circular
aperture. There is no smear, jitter or mirror aberrations. In this case the system MTF is simply

MTFsystem = ( MTFsampling )( MTFdiffraction )

where the diffraction MTF is been renormalized to the sampling cutoff frequency:
€ %
2 ' −1 ρ ρ % ρ (2 (
MTFdiffraction = cos − 1− ' * *
π' Q Q & Q ) *)
&

Recall that Q is the ratio of the sampling cutoff frequency to the diffraction cutoff frequency.
€ we compare three idealized systems, all having the same system cutoff
In what follows
frequency. In System 1 the sampling and diffraction cutoff frequencies are equal, so Q = 1. It
will serve as the reference system. In system 2, the diffraction cutoff frequency is twice that of
System 1 (say by doubling the aperture diameter) so for it Q = 1/2. In system 3, the sampling
frequency is twice that of System 1 (either by halving the pixel pitch or doubling the focal
length) so for it Q = 2.

Figure 30 shows the sampling, diffraction and product MTFs for the three systems. For
these idealized cases the product represents the system MTF. In all three charts the frequency is
normalized on the sampling cutoff frequency, which is the same value for Systems 1 and 2, but is
double for the third system. In absolute terms the three charts are lined up so they can be directly
compared. Comparing the product MTFs, the second system comes out the best in that its MTF
is the highest at all frequencies, while the first system is the worst. But this should not be
surprising, because the second system has twice the aperture diameter as the first, and the third
system has twice the sampling resolution as the first.

But does System 3 necessarily provide the best image quality? Image quality depends on
other factors besides MTF. One of these is signal to noise. But even here there is an advantage
for the lower Q system with the same integration time (but more on this later). But there is a note
of caution when it comes to aliasing. Indeed for a Q = ½ system aliasing occurs at all
frequencies, whereas for a Q = 2 system there is no aliasing.

28 R. R. Auelmann
11/5/12

Q=1

Q = 1/2

Q=2

Figure 30. MTFs for Three Idealized Systems

29 R. R. Auelmann
11/5/12

Fiete (Reference 2) has compared the imagery for cases closely related to the three
systems described in Figure 30. The only difference is that image sharpening has been applied,
which affects the MTF. Equal signal to noise ratios were assumed. The images (Figure 31) are all
viewed from directly overhead, so the ground sample distance (GSD) relates directly to the
sampling cutoff frequency. The objects to the side of each image represent the relative pixel and
aperture sizes. As expected, the second and third images are sharper than the first. Viewing the
bar charts, the predicted aliasing in the second image is quite apparent, while there is no aliasing
in the third image as expected.

Figure 31. Imagery Comparisons. Fiete, R. E., “Image Quality and λFN/p for Remote Sensing
Systems”, Optical Engineering, 38 (7), July 1999

It can be argued that in order to achieve the same signal to noise, the Q = 2 system would
require four times longer integration times than the Q = 1 system, and 16-times longer than the Q
= 1/2 system, and that the longer integration times make the Q = 2 system more sensitive to

30 R. R. Auelmann
11/5/12

image smear. This means that in order to achieve the full sharpness benefits of Q = 2 imagery,
one must pay much greater attention to limiting the amount of smear.

Edge Response Function.

Edge response (ER) is a measure of how well an imager is able to reproduce sharp edges
in a scene. The scene edge has zero value to the left of center, and unity value to the right of
center. It is defined by the equation

1 !/Q % MTF ( χ ) (
ER(ξ ) = 0.5 +
π
∫ 0
'
& χ
sin2πχξ *dχ
)

where ξ is the number of pixels offset measured from the edge, and χ is the spatial frequency
normalized on the sampling cutoff frequency 1/IFOV. Note that the upper limit of the integral is

the ratio of the optical cutoff frequency to the sampling cutoff frequency, or 1/Q. The MTF
includes all contributions (diffraction, sampling, smear, jitter, wavefront, boost, etc.). Figure 32
is a generic representation of the ER function. Differentiation of the ER with respect to the pixel
offset defines the Line Spread Function (LSF) closely resembling the PSF.

In most cases the LSF is a symmetric function about the zero offset point. But there are
exceptions arising from certain wavefront aberrations. Those are the cases when the OTF rather
than the MTF correctly models the system.

The relative edge response (RER) and the overshot parameter H are key parameters,
derived from the ER, that are used to assess image quality. The RER is the difference in ER
measured at a half a pixel on both sides the edge and is given by

RER = ER(0.5) − ER(−0.5)


1/Q % sin πχ (
= 2 ∫ 0 ' MTF ( χ ) *dχ
& πχ )

The overshoot parameter H is also determined from the ER function. It is the peak value of the
ER over the range ξ = 1 to 3 . However, if the ER monotonically increases (over this range), H is
the value of the ER at €
ξ = 1.25 .


31 R. R. Auelmann
11/5/12

ER(ξ)

H = ER(ξ = 1.25)

RER = ER(0.5) - ER(-0.5) 0.5

-3 -2 -1 0 1 2 3
Pixel displacement from edge, ξ

Figure 32. Edge Response (ER) and the Relative Edge Response (RER)

Figure 33 shows the ER for three ideal frame camera systems defined entirely by the
sampling and diffraction MTFs. The lower the optical Q the closer the ER follows the step input.
This because the diffraction spread is less at lower Q. The values of RER and H for ideal Q =
0.5, 1 and 2 systems are listed in Table 4. The manner in which the RER and H enter into the
assessment of image quality is discussed later. The corresponding LSFs are shown in Figure 34.

Figure 33. Edge Response ER for Ideal Diffraction Limited Q = 0.5, 1 and 2 Systems

32 R. R. Auelmann
11/5/12

TABLE 4. RER and H for Ideal Systems

Q RER H
0.5 0.747 0.971
1 0.600 0.946
2 0.383 0.878

Figure 34. Line Spread Functions for Ideal Systems

For scanning systems, the MTFs in the along scan x and cross scan y directions are
necessarily different. In such cases both the RER and H will be axis dependent. Such axis
dependence can also result from certain wavefront aberrations. In such instances both RER and
H are computed as the geometric means of their x and y components: RER = RERx ⋅ RERy and
H = H x ⋅ H y . Table 5 lists the RER and H values for ideal scanning systems for different Q.
These values can be compared to those in Table 1 an ideal non-scanning system. The results
differ only in the fact that the MTF for scanning has been included
€ in computing the x-axis
values for ER and H. The differences between the scanning and non-scanning values decrease as
€ Q increases, as expected.

TABLE 5. RER and H for Ideal “Scanning” Systems.

Q RERx RERy RER Hx Hy H


0.5 0.646 0.747 0.695 0.971 0.971 0.971
1 0.539 0.600 0.569 0.939 0.946 0.942
2 0.366 0.383 0.374 0.868 0.878 0.873

33 R. R. Auelmann
11/5/12

Edge Sharpening (MTFC). An image analyst normally has a “soft copy” of the image to
which he can apply different forms and amounts of edge sharpening to maximize his ability to
extract the information of interest from the image. The amount of sharpening that he can apply
largely depends on the signal to noise ratio (SNR) for that image. The higher the SNR the more
aggressively he can apply sharpening. The sharpening can be represented by its MTF. Such
sharpening is customarily denoted as MTFC where the C stands for compensation.

Unlike the system analyst, the system designer is faced with the tasked of predicting the
overall image quality ahead of time. To do so he must include a representative MTFC in the
overall system MTF when computing the system edge response ER. Unlike the other MTF
components, it affects image quality in two ways: through the relative edge response (RER) that
is based on the total system MTF, and through the amplification of noise, which has the same
effect as reducing the signal to noise ratio (SNR). This will be explained shortly.

There are a variety of sharpening algorithms. Perhaps the simplest is the 3 x 3 kernal

MTFC ( u,v ) = a + 2b(cos2πu + cos2πv) + 4c (cos2πu)(cos2πv )


MTFC (0,0) = 1
"c b c%
$ '
Kernel = $b a b' , a + 4b + 4c = 1
€ $#c b c'&

where u and v are two-dimensional frequencies normalized on the sampling frequency.


2 2 2
€ gain is G = a + 4b + 4c . Three examples are given in Table 6.
The corresponding noise

TABLE 6. Low, Medium and High Gain 3 x 3 Kernels


€ a 2.707 5.00 9.512
b -0.3536 -0.70 -1.5081
c -0.0732 -0.30 -0.62095
a+4b+4c 1.00 1.00 1.00
G 2.80 5.23 10.06

For a slice along the v = 0 axis, the MTFC equation reduces to

MTFC ( u,0) = a + 2b(1+ cos2πu) + 4c (cos2πu)

This function is plotted in Figure 35 for the above gain examples. Note that the peak
amplification (3x for low gain and 12x for the high gain) occurs at the Nyquist frequency. Note
€ GIQE, is somewhat lower than the peak gain.
that G, used in the

34 R. R. Auelmann
11/5/12

Figure 35. MTFC Using 3 x 3 Kernel with 3 Gains

Figure 36 shows the application of low gain MTFC kernel to an ideal Q = 1 imaging
system, where the ideal MTF (before MTFC is applied) is defined by MTFsampling ⋅ MTFdiffraction .
The ideal MTF is in blue, the MTFC is in red and the product is in black. Values for RER and H
for the compensated system MTF are listed in the figure.

Figure 36. System MTF with Low Gain MTFC Kernel

35 R. R. Auelmann
11/5/12

The Wiener filter is another MTFC model that seeks to optimize performance on the
basis of the expected or estimated SNR denoted by S. It is given by

MTF
MTFC =
MTF 2 + 1/S 2

Note that the peak value of the MTFC is S/2 and it occurs when the uncompensated MTF = 1/S.
Figure 37 shows the application of a Wiener filter with S = 6, so chosen that it have the same
peak value of 3 as for the 3€x 3 Kernel example of Figure 34. Here too the uncompensated MTF
is for an ideal Q = 1 system. Both the MTFC and system MTF profiles are quite different than
those of Figure 34. In particular note that the Wiener filter is less than one for frequencies less
than 0.78 whereas the gain is always ≥ 1 for the 3 x 3 kernel. Also note that both the system RER
and H are lower for the Wiener filter. Since a high RER and a low H are preferred, one cannot
say which filter is better at this point. Unlike the previous MTFC model where the noise gain G
is computed from the kernel values, we propose using the peak gain value Gpeak = S/2, which is
conservative.
3.0

Wiener Filter MTFC


2.5 SNR = 6
Gpeak = 3
MTF, MTFC, MTF•MTFC

2.0

1.5

MTF•MTFC
1.0
RER = 0.890
H = 1.032

0.5

Ideal MTF for Q = 1


0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Frequency

Figure 37. System MTF with Low Gain Wiener Filter (S = 6)

The differences between the two MTFC filters are even more striking for a Q = 2 optical
system. The comparisons are shown in Figures 38 and 39 for the same low gain MTFC values.
Note that the frequency scale is normalized on the sampling frequency. In this case both the
system RER and H are now higher for the Wiener filter (a switch from the Q = 1 case).

36 R. R. Auelmann
11/5/12

3.0
3 x 3 MTFC
G = 2.8
Gpeak = 3
2.5

MTF, MTFC, MTF•MTFC

2.0

1.5

1.0
MTF•MTFC
RER = 0.548
H = 0.948
0.5

Ideal MTF for Q = 2

0.0
0.0 0.1 0.2 0.3 0.4 0.5
Sample Frequency

Figure 38. Low Gain MTFC Kernel Applied to Ideal Q = 2 System.

3.0

Wiener Filter MTFC


2.5 SNR = 6
Gpeak = 3
MTF, MTFC, MTF•MTFC

2.0

1.5

1.0 MTF•MTFC
RER = 0.595
H = 1.028

0.5

Ideal MTF for Q = 2

0.0
0.0 0.1 0.2 0.3 0.4 0.5
Sample Frequency

Figure 39. Low Gain Wiener Filter Applied to Ideal Q = 2 System.

To gain a feel for the impact of edge sharpening on image quality consider the simulated
imagery shown in Figures 40 and 41 in which a high resolution aerial photograph of Washington
National Airport was down-sampled using a filled circular aperture, Q = 2 sensor. The two

37 R. R. Auelmann
11/5/12

images in Figure 38 have no edge sharpening. At an SNR of 100 the noise is indiscernible, while
for an SNR of 10 it becomes quite evident. In Figure 39 edge sharpening (of an unspecified type)
is applied to the same two SNR images. The benefit of sharpening is evident when comparing
the high SNR image in Figures 31 to that in Figure 32. However, sharpening also amplifies the
noise as seen by comparing the low SNR cases in these same two figures. Indeed the
unsharpened image is preferred at low SNR.

SNR = 100 SNR = 10

Figure 40 Effects of Noise-No Sharpening (Rehfield)

SNR = 100 SNR = 10

Figure 41. Effects of Noise-With Sharpening (Rehfield)

38 R. R. Auelmann
11/5/12

Radiometry and Signal to Noise Ratio.

Image quality depends not only on the resolution and the edge response, but also on the
signal to noise ratio (SNR) where the signal is a proportional to the number of photons from the
target and the noise is due to a variety of sources. Here we examine the signal to noise ratio for
solar illuminated scenes with sensing in a reflective band. Typically these are the visible band
(0.4 µm – 0.7 µm), the VNIR band (0.4 µm – 0.9 µm) and the SWIR band (0.9 µm – 1.7 µm).
Determination of the SNR is complex requiring the use of a sophisticated radiometry model of
the atmosphere as a function of wavelength, seeing conditions, and locations of the Sun and
satellite relative to the target area to be imaged.

Solar Irradiance. The global irradiance (W/m2) is the total solar irradiance on a
horizontal surface of the Earth. It has two components:

E global = E direct ⋅ cosψ sun + E diffuse

where E direct is the direct irradiance due to the solar rays along a straight path between the Sun
and a surface normal to the Sun, E diffuse is the diffuse irradiance due to the indirect scattered solar

rays off the atmosphere to a locally horizontal surface, and ψ sun is the Sun zenith angle. E direct
secψ
€ includes the atmospheric transmission loss (τ atm )
sun
along the Sun-to-scene path.
E direct ⋅ cosψ sun is the component
€ of the direct solar irradiance on a horizontal surface. The direct
and diffuse irradiances are depicted in Figure 42. € €

Figure 42. Irradiance on a Horizontal Surface

To gain an appreciation of the complex interaction with cloud cover, consider the
daylight irradiances (Figure 43 and 44) measured at Golden Colorado on a clear day April 9,
2003 and a partly cloudy day July 3, 2004. The data is from the National Renewable Energy
Laboratory (NREL) and reported by Stoffel and Wilcox (see reference). The direct beam
radiances were measured using a pyrheliometer on a Sun following tracker. The pyrheliometer is
a narrow field of view sensor (≤ 5°) that measures the irradiance in the band 0.2 µm to 4 µm.

39 R. R. Auelmann
11/5/12

The global irradiances were measured using a horizontal pyranometer that senses the radiation
over a hemispherical FOV in the 3.1 µm to 28 µm band. The diffuse sky irradiances were
measured using a horizontal pyranometer with a shield on a Sun-following tracker to block the
direct solar rays.

Figure 43. Clear Sky Irradiances

Figure 44. Partly Cloudy Sky Irradiances

Because the direct beam irradiance is measured normal to the Sun line it can exceed the
global irradiance, which is measured on a horizontal surface. Note that the diffuse irradiance
typically increases with cloud cover and the increase can be significant.

40 R. R. Auelmann
11/5/12

Solar Radiance Components. As depicted in Figure 45 the total solar spectral radiance
at the top of the atmosphere in the direction of the sensor has three components:

Lλ = Ldirectλ + Ldiffuseλ + L pathλ (W ⋅ m −2


⋅ sr−1 ⋅ µm−1 )

where Ldirectλ is the direct spectral radiance due to solar rays that directly reflect off the scene,
Ldiffuseλ is the diffuse spectral radiance due to solar rays that bounce off aerosols (in the

atmosphere) and then reflect off the scene, and L pathλ is the path spectral radiance due to solar
€ rays that bounce off aerosols directly toward the sensor. Only Ldirectλ and Ldiffuseλ contribute to the
€ “scene signal”. L pathλ (or “haze” radiance) is strictly a noise component. Note it is the diffuse
component that allows one to image the € shadowed portions of the scene.
€ €

Figure 45. Three Radiance Components

The spectral radiance components reflected off the scene are related to the scene spectral
irradiances by the equations

ρ secψ
Ldirectλ = ⋅ E directλ ⋅ cosψ sun ⋅ τ atm
π
ρ secψ
(W ⋅ m −2
⋅ µm−1 ⋅ sr −1 )
Ldiffuseλ = ⋅ E diffuseλ ⋅ τ atm
π

where ρ is the scene reflectance. Assuming the surface


€ acts as a Lambertian reflector the
radiance is expressed per π steradians. Because the radiances are measured at the top of the

€ 41 R. R. Auelmann

11/5/12

secψ
atmosphere they include the atmospheric transmission loss (τ atm ) along the sensor LOS at the
sensor zenith angle ψ .

Codes such as MODTRAN and MOSART compute Ldirectλ , Ldiffuseλ and Lhazeλ for various

standard atmospheric models and aerosol models (expressed in terms of visibility) with the sun

and sensor zenith angles as inputs. For the most part, the diffuse and path radiances are
dependent on the aerosol content of the atmosphere.

Sensor Electron Flux. Because the sensors employ photon detectors it is more useful to
express the reflected scene radiance and the haze (or path) radiance in terms of photon flux
(photons/s) rather than power (W):

λ
RSλ =
hc
⋅ ( Ldirectλ + Ldiffuseλ ) ( photons ⋅ s
−1
⋅ m−2 ⋅ sr −1 ⋅ µm−1 )

λ
R Hλ =
hc
⋅ L pathλ ( photons ⋅ s −1
⋅ m−2 ⋅ sr−1 ⋅ µm−1 )

where h is Planck’s constant ( 6.6256 ×10−34 W ⋅ s2 ) and c is the speed of light ( 3 ×10 8 m /s ). Then
the electron flux output of the detector due to solar radiance is given by

2
2 π % D( λ2
Φ = (r ⋅ IFOV)€ ⋅ ' * ⋅ η fill ⋅ ∫ ηQE ( λ ) ⋅ τ opt ( λ) ⋅ [ RSλ + R Hλ ] ⋅€dλ (electrons /s)
4& r) λ1

2
where (r ⋅ IFOV) is the area of the scene normal to the LOS subtended by a square detector,
2
€ π # D & is the solid angle (sr) subtended by a filled circular aperture, η is the aperture fill
% ( fill
4$ r'
€ factor, ηQE is the detector quantum efficiency and τ opt is the optical train transmission. In terms
of the previously defined optical Q, the equation for Φ takes on the more instructive form

€ 2
π %λ ( λ2
€ Φ = ' mean * ⋅ η fill ⋅ ∫ λ ηQE ( λ€) ⋅ τ opt ( λ ) ⋅ [ RSλ + R Hλ ] ⋅ dλ (electrons /s)
4& Q ) 1 €

In particular, we see that Φ is independent of the slant range, and is the same for sensor
combinations of IFOV and D that have the same Q and mean wavelength.

In many instances ηQE and τ opt can be treated as constants over the spectral band of

interest in which case
Φ = K ⋅ ( RS + R H ) (electrons /s)
€ €
λ2 λ2
RS = ∫ λ1
RSλ ⋅ dλ, RH = ∫ λ1
R Hλ ⋅ dλ ( photons ⋅ s −1
⋅ m−2 ⋅ sr−1 )
where

€ 42 R. R. Auelmann
11/5/12

2
π $ λmean '
K= & ) η fill ⋅ ηQE ⋅ τ opt (electrons ⋅ photons−1 ⋅ m 2 ⋅ sr)
4% Q (

is the sensor “responsivity”.


€Scene and Haze Radiances. MOSART is an Air Force developed code for computing
radiances. Brandon Ho of Northrop Grumman used MOSART to compute the signal and path
radiance integrals for a Mid-Latitude Summer atmosphere with 23-km visibility for selected
combinations of Sun and sensor zenith angles. The vertical atmospheric transmission τ atm over
the spectral band 0.4 µm to 1.7 µm is shown in Figure 46. A scene reflectance ρ = 1 was
assumed so that ρ can be reassigned without rerunning the code.

1 €

0.8
Atmospheric Transmission

0.6

0.4

0.2

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
Wavelength, λ (µm)

Figure 46. Vertical Atmospheric Transmission

Figures 47 shows the corresponding total scene and haze spectral radiances at the top of
the atmosphere for the case when both the Sun and sensor are directly overhead. The direct and
diffuse components of the scene spectral radiance are also shown.

43 R. R. Auelmann
11/5/12

Figure 47. Reflected and Haze Spectral Radiances (Top of Atmosphere)

MOSART was used to compute VNIR and SWIR scene and path radiances for the cases
listed in Table 7 for the same atmospheric conditions used above. The first run series is for ψ sun
= 75° and ψ between 0 and 70°. A second run series is for ψ = 40° and ψ sun from 0 to 90°. The
code output the two integrals RS = Rdirect + Rdiffuse (for a reflectance ρ = 1) and R H for each
spectral band and for each combination of Sun and sensor zenith angles. The results € are listed in
Table 8. The direct radiance component goes to zero for a ψ sun ≥ 90° (as expected). However,
€ € €
the code gives erroneous results for the diffuse component under twilight conditions when ψ sun >
€ €
90. MODTRAN has€the same problem.

TABLE 7. MOSART Runs

Sensor Zenith
Sun Zenith 0 40 60 70
0
60
75
90

44 R. R. Auelmann
11/5/12

TABLE 8. Scene and Path Radiances (1018 ph /s /m 2 /sr)

Sun ψsun = 90°


VNIR (0.4 µm - 0/9 µm) SWIR (0.9 µm - 1.7 µm)
Sensor ψ Rdirectcosψsun Rdiffuse RS RH Rdirectcosψsun Rdiffuse RS RH
40 0 95.1 95.1

22 0 105 105 6.4
60 0 78.1 78.1 30.1 0 96.8 96.8 9.1

Sensor ψ = 0°
VNIR (0.4 µm - 0/9 µm) SWIR (0.9 µm - 1.7 µm)
Sun ψsun Rdirectcosψsun Rdiffuse RS RH Rdirectcosψsun Rdiffuse RS RH
0 375 184 559 54.7 359 63 422 7

Sensor ψ = 40°
VNIR (0.4 µm - 0/9 µm) SWIR (0.9 µm - 1.7 µm)
Sun ψsun Rdirectcosψsun Rdiffuse RS RH Rdirectcosψsun Rdiffuse RS RH
0 340 165 505 63.7 341 60 401 8.8
60 124 154 278 56.5 146 57.1 203.1 7.6
75 37.5 173.5 211 53.3 58.8 68.2 127 7.1
90 0 95.1 95.1 22 0 105 105 6.4

Sun ψsun = 75°


VNIR (0.4 µm - 0/9 µm) SWIR (0.9 µm - 1.7 µm)
Sensor ψ Rdirectcosψsun Rdiffuse RS RH Rdirectcosψsun Rdiffuse RS RH
0 40.9 193.1 234.0 40.3 61.2 71.8 133.0 5
40 37.5 173.5 211.0 53.3 58.8 68.2 127.0 7.1
60 31.0 136.0 167.0 74.0 53.9 62.1 116.0 10.7
70 24.3 100.7 125.0 93.0 48.2 55.8 104.0 14.9

The two run series are graphed in Figures 7 and 8, each for the VNIR and SWIR.
Recognize that the direct radiance goes to zero if a cloud interferes with the Sun path to the
scene, in which case the “signal” is due solely to the diffuse radiance. And the presence of clouds
tends to increase the magnitude of the diffuse component. The question is logically raised
whether either band has a SNR advantage over the other, under stressing conditions.

Figure 48. VNIR and SWIR Radiances for Range of Sun Zenith Angles for ψ = 40°

45 R. R. Auelmann

11/5/12

Figure 49. Radiances for Range of Sensor Zenith Angles for ψ sun = 75°

Noise Sources. Photon noise is due to random fluctuations in the arrival rate of photons
from both the reflected scene radiance and the path radiance. Because photon noise follows a
Poisson distribution, the standard deviation is equal to the square€root of the expected value (the
mean) we have
σ photon = K ( ρ ⋅ RS + R H ) ⋅ t

Even in the absence of an input signal the detector will produce an output due to dark
current. The dark current noise also follows a Poison distribution equal to the square root of the
product of the pixel dark €
current ΦD (electrons/sec) and the integration time tint.

σ dark = ΦD ⋅ t

Typically it can be reduced to small values with cooling. Readout noise σ readout is an inherent
property of a detector that occurs when amplifying and converting photoelectrons into a voltage.
Typical readout noise values for€CCD arrays are between 5 and 50 noise electrons.
€ counts using an analog to
The analog output of the detector is converted into digital
digital converter. The well size N well is the maximum number of electrons that can be stored in a
single detector element. To provide b bits resolution, N well must be divided into 2 b digital
counts. Thus N well /2 b is the number of electrons per digital count. The corresponding noise is

1 N well €
€ =
σ quantization ⋅
12 2 b

€ 46 R. R. Auelmann
11/5/12

Signal to Noise Ratio. To compute the target signal to noise ratio (SNR) one must
assume values for the target reflectance. For imaging in the VNIR and SWIR bands the 1996
General Image Quality Equation (GIQE) Users Guide recommends a delta reflectance Δρ = 0.08
to compute the target signal, and a reflectance ρ = 0.15 to compute the random noise due to the
target signal. Accordingly, the SNR is given by

Δρ ⋅ K ⋅ RS ⋅ t int €
SNR = €
[K (ρ ⋅ RS + RH ) + ΦD ] ⋅ t int + σ readout
2 2
+ σ quantization

As the integration time increases, the readout and quantization errors become negligible,
in which case the SNR reduces to

Δρ ⋅ K ⋅ RS
SNR = ⋅ t int
[ ( S H ) D]
K ρ ⋅ R + R + Φ

This holds true even for a readout noise of 50 electrons and an integration time of 0.001 sec. And
when K ( ρ ⋅ RS + R H ) >> N D , which is nearly always the case, the SNR can be written

Δρ ⋅ RS
SNR = ⋅ K ⋅ t int
€ ρ ⋅ RS + R H

It follows that the SNR scales as:



SNR ∝ ( λmean /Q) ⋅ η fill ⋅ ηQE ⋅ τ opt .

Consider now the evaluation of the above SNR equation for the VNIR and SWIR bands
for the following conditions: ψ = 40 with ψ sun = 0° to 90°. Assume the default values for
reflectance, Q = 1, η€fill = 1, ηQE = 0.7, τ opt = 0.7 for which

−13 −1 2
€ VNIR : K€= 1.63 ×10 electrons ⋅ photons ⋅ m ⋅ sr
€ SWIR : K = 6.52 ×10−13 electrons ⋅ photons−1 ⋅ m 2 ⋅ sr

The SNR is plotted in Figure 50 for the two bands with and without direct radiance for a 1-ms
integration time. Direct radiance equal to zero corresponds to having the direct rays of the Sun
blocked by a€ cloud. Under these conditions the SNR for the SWIR appears roughly a factor of
two higher than for the VNIR. The SNRs would be roughly equal if the SWIR Q were twice that
of the VNIR Q.

47 R. R. Auelmann
11/5/12

Figure 50. SNR for VNIR and SWIR with and without Direct Radiance

48 R. R. Auelmann
11/5/12

Image Quality Model.

Image quality is customarily defined in terms of NIIRS, which stands for the National
Imagery Interpretability Rating Scale. It is a ten-level scale that allows image analysts to rate the
usefulness of an image to perform a specific task, ranging from the easiest – Level 0 (e.g. the
detection and identification of large scale objects) to the most difficult – Level 9 (identification
and analysis of small objects or features). The NIIRS-scale was developed in the 1970s by the
US Government using visible imagery to perform intelligence tasks involving mainly military
related objects. Over the years, separate NIIRS scales have been developed for infrared imagery
applied to intelligence applications, and visible imagery applied to other fields of interest
including civil, agriculture, cultural, and natural.

GIQE. The General Image Quality Equation (GIQE) is an empirical tool that predicts the
NIIRS performance of an optical system under a given set of image collection conditions
(Leachtenauer, et al. Applied Optics Nov 1997). While the GIQE was originally developed in the
1980s, it was not formally released as Version 3 to the public until 1994. Its attractiveness lies in
the fact that it does a reasonably accurate job of predicting NIIRS using only five engineering
parameters: the ground sample distance (GSD), the relative edge response (RER), the geometric
height overshoot due to edge sharpening (H), the signal to noise ratio (SNR), and the noise gain
due to edge sharpening (G). The latest (Version 4) was introduced in 1996. Both Versions 3 and
4 have the same form:

NIIRS= c0 + c1 ⋅ log10 GSD + c2 ⋅ log10 RER + c3 ⋅ H + c4 ⋅ G/SNR

However, the coefficients differ as shown in Table 9. Also there are two sets of coefficients for
Version 4 depending on the value of the RER.

TABLE 9. GIQE Coefficients (English Units for GSD)

GSD (inches) GSD RER H G/SNR


co c1 c2 c3 c4
GIQE 3 11.81 -3.32 3.32 -1.48 -1
GIQE 4 (RER < 0.9) 10.251 -3.16 2.817 -0.656 -0.344
GIQE 4 (RER ≥ 0.9) 10.251 -3.32 1.559 -0.656 -0.344

Because these notes use metric values throughout, I have converted this table so that the GSD is
expressed in meters, which results in changes to co as shown in Table 10.

TABLE 10. GIQE Coefficients (Metric Units for GSD)

GSD (meters) GSD RER H G/SNR


co c1 c2 c3 c4
GIQE 3 6.555 -3.32 3.32 -1.48 -1
GIQE 4 (RER < 0.9) 5.205 -3.16 2.817 -0.656 -0.344
GIQE 4 (RER ≥ 0.9) 4.955 -3.32 1.559 -0.656 -0.344

49 R. R. Auelmann
11/5/12

Note the significance of the coefficient value 3.32: with it a factor of two change in the
parameter (GSD or RER) represents delta-NIIRS of ± 1 depending on the sign of the coefficient;
that is, 3.32log10 2 = 1, which seems intuitively correct. For example, a factor of two in distance
should result in a change of one NIIRS. In this respect the Version 3 “form” seems more
intuitively correct than Version 4. Indeed, Thurman and Fienup (“Analysis of the General Image
Quality Equation”, Proc of the SPIE, vol 6978 (2008)) make just this point and present measured
€ results that are pertinent.

The transition point (at RER = 0.9) is also pause for concern for Version 4. Consider the
example in Table 11 that is computed for: GSD = 1-m, RER = 0.9, H = 1.033 and a range of
SNR. It is the result of an “ideal” Q = 1 system with a Wiener filter applied with S = 6.28. The
delta NIIRS is 0.19 between Versions 4.0a and 4.0b, which is excessive.

TABLE 11. Example of GIQE Version 4 Differences at the Transition Point RER = 0.9

Q = 1 Wiener filter NIIRS for 1-m GSD


SNR SNR SNR SNR SNR SNR SNR
Version S RER H Gpeak 1 2 5 10 20 50 100
4.0a 6.28 0.9 1.033 3.14 3.32 3.86 4.18 4.29 4.34 4.38 4.39
4.0b 6.28 0.9 1.033 3.14 3.13 3.67 3.99 4.10 4.15 4.18 4.20

It is also important to understand the limitations under which the GIQE 4 was developed.
The evaluated imagery was for a Q = 1 optical system with the following range of values:

GSD 3 in to 80 in (0.08 m to 2 m)
RER 0.2 to 1.3
SNR 2 to 130
G/SNR 0.01 to1.8
H 0.9 to 1.9

Values outside this range may or may not be valid.

Consider now the three systems represented in Figure 31 and compute their NIIRS on the
assumption that these are ideal systems (characterized by their sampling and diffraction limited
MTFs). We further assume no edge sharpening and a high SNR. Results are shown in Table 12.
Note that the Q = 0.5 and Q = 2 results are beyond the test data for Version 4.

TABLE 12. NIIRS Predictions for an Ideal Frame Camera

Q GSD RER H G SNR NIIRS


0.5 1 0.747 0.971 1 50 4.21
1 1 0.600 0.946 1 50 3.96
2 0.5 0.383 0.878 1 50 4.40

The corresponding results on the assumption of an ideal scanner are given in Table 13. The sole
difference between these two tables is the inclusion of scan smear. The biggest difference
between the frame camera and the scanner is only a tenth of a NIIRS occurring for Q = 0.5 case.

50 R. R. Auelmann
11/5/12

TABLE 13. NIIRS Predictions for an Ideal Scanner

Q GSD RER H G SNR NIIRS


0.5 1 0.695 0.971 1 50 4.12
1 1 0.569 0.942 1 50 3.90
2 0.5 0.374 0.873 1 50 4.38

The Q = 1 and Q = 2 cases have the same aperture diameter, so they differ only in that the
Q = 2 system has double the sampling. The predicted differences in NIIRS for these two systems
are 0.44 for the frame camera and 0.48 for the scanner. This advantage is somewhat misleading
in that we have assumed the same SNR. However, a Q = 2 system requires four times longer
integration time to achieve the same SNR as a Q = 1 system, making it more sensitive to smear
and jitter. Some of this advantage can be expected to vanish when these effects are included.

Next we compute NIIRS for an “ideal” Q = 1 with a 1-m GSD and different MTFC filter
models using Version 4. Table 14 shows values for a 3 x 3 Kernel with different gains as a
function of SNR. Note that the two high gain values for RER and H using the 3 x 3 kernel
exceed the range of measured values listed above. Table 15 shows corresponding results using a
Wiener filter with different values of S. Both tables include the zero gain option. The Wiener
filter appears to provide somewhat better results than the 3 x 3 kernel. And At high SNR the
Wiener filter predicts an improvement of up to 0.36 NIIRS, which is substantial.

TABLE 14. NIIRS Predictions for Ideal Q = 1 System Using 3 x 3 Kernel MTFC

Q=1 NIIRS for 1-m GSD


SNR SNR SNR SNR SNR SNR SNR
G RER H Gpeak 1 2 5 10 20 50 100
1 0.600 0.946 3.62 3.79 3.89 3.93 3.94 3.95 3.96
2.8 0.952 1.145 3 3.21 3.69 3.98 4.07 4.12 4.15 4.16
5.23 1.575 1.479 6.2 2.49 3.39 3.93 4.11 4.20 4.26 4.27
10.06 2.69 2.084 12 0.80 2.53 3.57 3.91 4.08 4.19 4.22

TABLE 15. NIIRS Predictions for Ideal Q = 1 System Using Wiener MTFC

Q=1 NIIRS for 1-m GSD


SNR SNR SNR SNR SNR SNR SNR
S RER H Gpeak 1 2 5 10 20 50 100
NA 0.6 0.946 NA 3.62 3.79 3.89 3.93 3.94 3.95 3.96
1 0.261 0.722 0.5 2.92 3.00 3.05 3.07 3.08 3.08 3.09
2 0.532 0.899 1 3.50 3.67 3.77 3.81 3.83 3.84 3.84
5 0.845 1.023 2.5 3.47 3.90 4.16 4.24 4.28 4.31 4.32
10 0.988 1.061 5 2.53 3.39 3.91 4.08 4.16 4.22 4.23
20 1.071 1.074 10 0.86 2.58 3.61 3.95 4.12 4.23 4.26

51 R. R. Auelmann
11/5/12

Double Sampling. The GSD, RER and H are computed as geometric mean values along
x and y axes that are aligned to the sampling axes of the detector arrays. These axes may not be
aligned to either the scan direction or the satellite track. Double sampling is a good example. It
involves combining two offset image frames to form a single frame with a √2 higher sampling
resolution. The French Supermode image scheme, discussed earlier, is an example of this
approach. The French claim (incorrectly I contend) a factor of two-gain in sampling resolution.

Consider a frame camera with an array of N x N pixels each with a side dimension d.
Normally the pixel pitch (spacing between pixels) p is equal to d, in which case the IFOV is
simply d/f where f is the focal length. By double sampling it is possible to reduce the sampling
distance and the effective IFOV by the √2 without changing the focal length. To accomplish this
the two images must be shifted a distance p = d/√2 at a 45° angle to the major array axes as
depicted in Figure 51. The two frames (red and blue) are shown on the left and the corresponding
sample positions are represented by the dots. The synthesized pixel array (in black) are centered
on the sample points and are aligned at 45° to the red and blue arrays. It has √2 times more pixels
than the either of the original red or blue frames, and each synthesized pixel area is √2-times
smaller than the area of the sensor pixels.

Figure 51. Double Sample Geometry

A single synthesized pixel is represented in Figure 52. Its value c11 is expressed as a
function of the six shaded blue aij and red pixel bij values. Note that a11 and b22 do not contribute
to c11. Also note that if the image were of uniform intensity in the vicinity of the synthesized
pixel, the synthesized pixel signal would be the same as either the red and blue pixel signals, but
over an area the √2-times smaller. This suggests that the SNR for the synthesized frame is equal
to the SNR of an individual red or blue frame.

52 R. R. Auelmann
11/5/12

Figure 52. Computation of the Elements of the Interlaced Frame

This makes sense when one considers that the optical Q for the synthesized frame is √2
times larger than the optical Q for either of the constituent frames. For a shot-noise limited
design the SNR is inversely proportional to Q squared. This would suggest that the SNR for the
synthesized frame would be only half that of a constituent frame. But the deficit is made up with
two frames.

53 R. R. Auelmann
11/5/12

Pan Sharpened Multi-Color Imaging

Most current commercial imaging satellites employ line scanners with a pan band and
lower resolution color bands. Consider the case a pan band and three color (red, green and blue)
bands each at four times lower resolution than the pan band. Also assume that the pan band has
four detectors in TDI, so that the readout rate is four times faster than each of the color bands,
and the integration time is the same for all four bands. The detector geometry is depicted in
Figure 53.

Figure 53. Pan Ban with Three Color Bands

Pan sharpening refers to a technique where the higher resolution pan image is used to
sharpen the apparent resolution of the color bands. Alternately, one can describe it as coloring
the higher resolution pan imagery. Intensity-Hue-Saturation (IHS) is one technique for pan
sharpening. In its simplest form it involves: first, transforming the red (R), green (G) and blue
(B) intensities to a multi-color intensity (I), hue (H) and saturation (S); second, replacing I by the
pan intensity (P); and third, transforming back to fused colors FR, FG and FB on a grid
corresponding the pan resolution.

First four image grids corresponding to the pan band pixel resolution are set up: one grid
for the pan band and one for each of the three colors. The intensities are entered on the grid. For
each color band the entries are 1/16 of the value of the corresponding color pixel. Suppose all
three colored intensities R, G and B were equal. Then on average we would expect each of the
color entries to be 1/3 of the pan intensity. The mathematical description of the IHS method is as
follows

I # 13 1
3
1
3
& R
% (
Step 1: H = %− 62 − 6
2 2 2
6 (⋅ G

S %$ 12 − 1
2
0 (' B

Step 2: Replace I by P

54 R. R. Auelmann
11/5/12

FR #1 − 12 1
2
& P #1 − 1
2
1
2
& I 1 R+ P−I
% ( % (
Step 3: FG = %1 − 12 − ( ⋅ H = %1 − 12
1
2
− ( ⋅ H + ( P − I) ⋅ 1 = G + P − I
1
2
FB %$1 2 0 (' S %$1 2 0 (' S 1 B+ P−I

2R G B
FR = − − +P
€ 3 3 3
R 2G B
Finally we have FG = − + − +P
3 3 3
R G 2B
FB = − − + +P
3 3 3

Application of this technique to IKONOS imagery is depicted in Figure 54. At the left is the
image produced using the four-meter resolution RBG bands, in the center is the same scene

imaged by the 1-m resolution pan band and at the right is the pan-sharpened image. The
colorized pan-sharpened image appears to have the resolution of the pan band alone. Actually
there is some small loss in spatial resolution. I am unaware of any studies that quantify the
resolution difference.

4-m resolution RBG 1-m resolution Pan Fused RBG (using IHS)

Figure 54. IKONOS imagery example of IHS Pan Sharpening (Yun Zhang ref)

However, there is color distortion in the sharpened image. This is most evident in the
blues and greens when comparing the RGB and Fused RBG images. The main reason for this
distortion is the fact that the IKONOS pan band extends well beyond the visible spectrum to
include a near infrared band (NIR) in addition to the R, B and G bands. The IHS method can be
extended to include the NIR band in which case we would have

55 R. R. Auelmann
11/5/12

3R G B NIR
FR = − − − +P
4 4 4 4
R 3G B NIR
FG = − + − − +P
4 4 4 4
R G 3B NIR
FB = − − + − +P
4 4 4 4
R G B 3NIR
FNIR = − − − + +P
4 4 4 4

Figure 55 is indicative of the reduction in color distortion when using such more
sophisticated techniques. This is a natural color image in which only the fused RGB are
displayed. Other more€sophisticated techniques are described in the references.

Figure 55. Enhanced IHS Technique Applied to IKONOS Image (Yun Zhang ref)

56 R. R. Auelmann
11/5/12

References

These papers provide general background information:

Pease, C. B., Satellite Imaging Instrument, Ellis Horwood, New York 1990.

Higgins, George C., “Methods for Engineering Photographic Systems, Applied Optics, vol 3, no
1, January 1964. Shows use of Kodak empirical equation.

Vitale, T., “Film Grain, Resolution and Fundamental Film Particles”, tvitale@.netcom.com April
2007. Shows Fuji equation from Fuji Data Guide (1998). Also lists Kodak empirical equation.

Fiete, Robert D., “Image Quality and λFN/p for Remote Sensing Systems”, Optical Engineering,
vol 38, no 7, July 1999.

Fiete, Robert D., and Tantalo, Theodore, “Image Quality of Increased Along-Scan Sampling for
Remote Sensing Systems”, Optical Engineering, vol 38, no 5, May 1999.

Light, D. L., “Characteristics of Remote Sensors for Mapping and Earth Science Applications”,
Photogrammetric Engineering & Remote Sensing, vol 56, No 12, December 1990, pp. 1613-
1623.

These papers discuss radiometry and signal to noise:

Stoffel, Tom and Wilcox, Steve, “Solar Radiation Measurements” A Workshop, August 4, 2004

Fiete, Robert D., and Tantalo, Theodore, “Comparison of SNR Image Quality Metrics for
Remote Sensing”, Optical Engineering, vol 40, no 4, April 2000.

These papers discuss the image quality equation:

Leachtenauer, Jon C. et al, “General Image-Quality Equation: GIQE”, Applied Optics, vol 36, no
32, November 10, 1997.

Thurman, S. T. and Fienup, J. R., “Analysis of the General Image Quality Equation”, Proc of SPIE, vol
6978 (2008).

Thurman, S. T. and Fienup, J. R., “Wiener Filtering of Aliased Imagery”, Proc of SPIE, vol 7076 (2008).

These papers provide basis for Intensity-Hue-Saturation technique for pan sharpening:

M. J. Choi, et al, “An Improved Intensity-Hue-Saturation Method for IKONOS Image Fusion”, to be
published in International Journal of Remote Sensing.

Yun Zhang, “Problems in the Fusion of Commercial High-Resolution Satellite as well as Landsat
7 Images and Initial Solutions” ISPRS, IGU, CIG.

57 R. R. Auelmann

Das könnte Ihnen auch gefallen