The art of radiometry
The art of radiometry
Palmer, James M.
Art of radiometry / James M. Palmer and Barbara G. Grant.
p. cm. -- (Press monograph ; 184)
Includes bibliographical references and index.
ISBN 978-0-8194-7245-8
1. Radiation--Measurement. I. Grant, Barbara G. (Barbara Geri), 1957- II.
Title.
QD117.R3P35 2009
539.7'7--dc22
2009038491
Published by
SPIE
P.O. Box 10
Bellingham, Washington 98227-0010 USA
Phone: +1 360.676.3290
Fax: +1 360.647.1445
Email: Books@spie.org
Web: http://spie.org
The content of this book reflects the work and thought of the author(s).
Every effort has been made to publish reliable and accurate information herein,
but the publisher is not responsible for the validity of the information or for any
outcomes resulting from reliance thereon.
5.4.4 Noise factor, noise figure, and noise temperature ............... 143
5.4.5 Some noise examples ......................................................... 144
5.4.6 Computer simulation of Gaussian noise.............................. 147
5.5 Thermal Detectors ........................................................................... 147
5.5.1 Thermal circuit ..................................................................... 147
5.5.2 Thermoelectric detectors ..................................................... 150
5.5.2.1 Basic principles ............................................................ 150
5.5.2.2 Combinations and configurations ................................. 153
5.5.3 Thermoresistive detector: bolometer ................................... 155
5.5.4 Pyroelectric detectors .......................................................... 157
5.5.4.1 Basic principles ............................................................ 157
5.5.4.2 Pyroelectric materials ................................................... 160
5.5.4.3 Operational characteristics of pyroelectric detectors ... 162
5.5.4.4 Applications of pyroelectric detectors........................... 162
5.5.5 Other thermal detectors....................................................... 163
5.6 Photon Detectors ............................................................................. 164
5.6.1 Detector materials ............................................................... 164
5.6.2 Photoconductive detectors .................................................. 169
5.6.2.1 Basic principles ............................................................ 169
5.6.2.2 Noises in photoconductive detectors ........................... 173
5.6.2.3 Characteristics of photoconductive detectors .............. 174
5.6.2.4 Applications of photoconductive detectors ................... 175
5.6.3 Photoemissive detectors ..................................................... 175
5.6.3.1 Basic principles ............................................................ 175
5.6.3.2 Classes of emitters....................................................... 176
5.6.3.3 Dark current ................................................................. 181
5.6.3.4 Noises in photoemissive detectors ............................... 182
5.6.3.5 Photoemissive detector types ...................................... 183
5.6.4 Photovoltaic detectors ......................................................... 185
5.6.4.1 Basic principles ............................................................ 185
5.6.4.2 Responsivity and quantum efficiency ........................... 195
5.6.4.3 Noises in photovoltaic detectors .................................. 196
5.6.4.4 Photovoltaic detector materials and configurations ...... 198
5.7 Imaging Arrays ................................................................................ 199
5.7.1 Introduction.......................................................................... 199
5.7.2 Photographic film................................................................. 199
5.7.2.1 History .......................................................................... 199
5.7.2.2 Physical characteristics ................................................ 201
5.7.2.3 Spectral sensitivity ....................................................... 201
5.7.2.4 Radiometric calibration................................................. 201
5.7.2.5 Spatial resolution.......................................................... 202
5.7.2.6 Summary ...................................................................... 202
5.7.3 Electronic detector arrays.................................................... 203
5.7.3.1 History .......................................................................... 203
5.7.3.2 Device architecture description and tradeoffs .............. 203
Index / 361
In 2006, Jim Palmer was told that he was terminally ill, and he asked me to
complete this work. I was humbled and honored by the request. I’d met Jim as a
graduate student in optical sciences in the late 1980s, and he had served on my
thesis committee. My career after graduation had focused on systems engineering
and analysis, two areas in which radiometry plays a significant role. For nearly
the last ten years of Jim’s life, I’d been able to receive mentoring from the master
simply by showing up at Jim’s office door with a question or topic for discussion,
but I never anticipated that our discussions would one day come to an end. Upon
Jim’s death, I sought to weave his collection of resources and narrative together
xi
with newer material and discussion in a manner I hope will be both informative
to read and valuable to reference. The preface that follows was written by Jim
before he died and has been left as he wrote it.
I am grateful for the assistance of many. First is William L. Wolfe, Jim’s
professor and mentor, who offered helpful comments on each chapter and
adapted Chapter 6 on radiometric instrumentation. Others for whose help I am
grateful, all from or associated with the University of Arizona College of Optical
Sciences, are Bob Schowengerdt, who contributed the narrative on film; Anurag
Gupta of Optical Research Associates, Tucson, Arizona, who adapted the
appendix material; and L. Stephen Bell, Jim’s close friend and colleague, who
revised the signal processing discussion that appears in that section and provided
a complete bibliography on the subject. A special note of thanks goes to Eustace
Dereniak, who provided office space for me, helpful discussions, and hearty
doses of encouragement. I also wish to thank John Reagan, Kurt Thome (NASA
Goddard Spaceflight Center, Greenbelt, Maryland), Mike Nofziger, and Arvind
Marathay for review, discussion, and helpful insights. In addition, I am grateful
for the assistance of Anne Palmer, Jim’s beloved sister, and University of
Arizona College of Optical Sciences staff members Trish Pettijohn and Ashley
Bidegain. Gwen Weerts and Tim Lamkins of SPIE Press have my gratitude for
the special assistance they provided to this project. I also gratefully acknowledge
Philip N. Slater, my professor in optical sciences, who selected me as a graduate
student and trained me in remote sensing and absolute radiometric calibration
from 1987 to 1989, and Michael W. Munn, formerly Chief Scientist at Lockheed
Martin Corporation, who instilled the value of a systems perspective in the
approach to technical problems.
Finally, I am grateful to my family for providing financial support; to Ralph
Gonzales, Arizona Department of Transportation, and Sylvia Rogers Gibbons for
providing professional contacts; and my friends at Calvary Chapel, Tucson,
Arizona, whose donations and prayers sustained me as I worked to complete this
book.
Barbara G. Grant
Cupertino, California
October 2009
James M. Palmer
1937–2007
xiii
-metry [Gr. -metria < metron] a terminal combining form meaning the process,
art, or science of measuring
These definitions are taken from Webster's New World Dictionary, and may
be satisfactory for the general nonscientist. The definitions are not satisfactory,
however, for scientists and engineers pursuing the art of radiometry. So let’s get
technical:
Figure 1.1 Classic vane radiometer, commonly called the Crooke radiometer.1 [Reprinted
by permission from Webster’s Third New International® Dictionary, Unabridged ©1993 by
Merriam-Webster, Incorporated (www.Merriam-Webster.com)].
The optical radiation spectrum will be treated in this text, including the
ultraviolet, visible, and infrared regions. The visible portion of the optical
spectrum covers a rather narrow band of wavelengths between 380 nm and
760 nm; the radiation between these limits, perceivable by the unaided normal
human eye, is termed “light.” Measurements within this region may be called
“photometric” if the instruments used incorporate the response of the eye. The
short wavelength (ultraviolet) limit of radiometric coverage is about 200 nm,
approximately the shortest wavelength that our atmosphere will transmit. The
longest wavelength (infrared) treated in this book is about 100 μm. This
wavelength range includes 99% of the energy (95% of the photons) from a
thermal radiator at 0° C (273.16 K).
I often say that when you can measure what you are speaking
about, and express it in numbers, you know something about it;
but when you cannot measure it, when you cannot express it in
numbers, your knowledge is of a meager and unsatisfactory
kind; it may be the beginning of knowledge, but you have
scarcely, in your thoughts, advanced to the stage of science,
whatever the matter may be.
Lord Kelvin
If you are really doing optics, you get photons under your
fingernails.
James M. Palmer
Measurement is the point at which the rubber meets the road. Hypotheses,
uncorroborated by measurement, cannot fulfill the same function. And if rubber
doesn’t meet the road, the car cannot move.
The measurement of light is often critical in transitioning from theory to the
development of systems and techniques. Although instrument and system design
may be based on theory, performance evaluation and system improvement
require that accurate radiometric measurements be applied. When calibrated
measurements are needed, that is, when field or laboratory measurements must
be correlated with specific values presenting the relationship between measured
phenomena and an absolute standard, radiometric measurements take on even
greater significance.
polarizers, optical fibers, etc. The optical radiation transmitted through the
instrument is finally incident on one or an array of detectors, transducers which
convert the incident optical radiation to a more tractable form of energy.
Detectors may be thermal (thermoelectric, bolometric, and pyroelectric) or
photon (photoemissive, photoconductive, and photovoltaic) in character; other
viable detectors include the human eye and photographic film.
The final block in the diagram involves signal processing. Most of the
detectors in common use generate electrical signals. Postdetector processing may
include filtering, linearization, and background suppression before the processed
result is recorded and displayed. Even the eye and film detectors include
processing steps, such as the filtering and interpretation of information by the
brain and the photographic development process for film.
Appearance measurement
Astrophysics
Atmospheric physics
Clinical medicine
Colorimetry
Diagnostic medicine
Remote-sensing satellites
Electro-optics
Illumination engineering
Laser measurements
Materials science
Meteorology
Military systems
Night-vision devices
Optoelectronics
Photobiology
Photochemistry
Photometry
Radiative heat transfer
Solar energy
Television systems
Visual displays
Vision research
Most books on radiometry begin with a vast and often confusing array of
terms, definitions, etc. In this work, detailed listing of terminology is relegated to
the glossary in Appendix E, and radiometric terms will be introduced as they are
needed.
Radiometry and photometry are applied to a variety of phenomena whose
output occurs over many orders of magnitude. Tables 1.3 through 1.5 illustrate
the ranges of illumination encountered. “Luminance” is power per unit area and
unit solid angle weighted by the spectral response of the eye; its units are lumens
per square meter per steradian (lm/m2/sr), or candelas per square meter (cd/m2).
“Illuminance” is power per unit area weighted by the same function; its units are
lumens per meter squared (lm/m2).*
*
For more thorough discussions of photometry, see J. T. Walsh, Photometry, Dover, New York
(1958); P. Moon, The Scientific Basis of Illuminating Engineering, Dover, New York (1961); and
R. McCluney, Introduction to Radiometry and Photometry, Artech House, Boston (1994).
References
(a) (b)
Figure 2.1 (a) Plane and (b) diverging wavefronts, with arrows indicating the direction of
radiation propagation.
11
dA1 dA2
Figure 2.2 A beam between two area elements.
The speed of light is the rate at which light propagates through a vacuum. It
is represented by c and is a constant equal to about 3.00 × 108 ms–1, a faster rate
of travel than in any other medium. The index of refraction of a medium is the
ratio of the speed of light in vacuum to the speed of light in the medium:
c
n= , (2.1)
v
where
Note from Fig. 2.3 that the refracted ray is closer to the normal to the
boundary between media when the index of refraction is higher; that is, when
n2 > n1. Also note that a cone of light becomes narrower in a higher-index
medium.
θ1
θ2
n1=1 n2=2
Ap = cos θ dA . (2.3)
A
Some common examples of projected area are shown in Table 2.2. Figure 2.4
depicts the relationships between surface and projected area for a circle and a
sphere.
Plane angles and solid angles are both derived units in the SI system. Figure
2.5 depicts the plane angle θ, with l the arc length and r the circle radius. The
solid angle ω extends the plane-angle concept to three dimensions. It is the ratio
of the element of spherical area dAsph to the square of the sphere radius r2. Figure
2.6 illustrates a solid angle.
The unit of the plane angle is the radian, defined as:
The radian is the plane angle between two radii of a circle that cut off on
the circumference an arc equal in length to the radius.1
Table 2.2 Shapes and corresponding surface- and projected-area formulas.
θ Ap (a)
Asph (b)
Ap
Figure 2.4 Surface- and projected-area relationships for a (a) circle and (b) sphere.
θ θ = l/r
dAsph
r
ω = dAsph
r2
rsinθdφ
dφ rdθ
dAsph
dθ
θ
φ
rsinθdφ
rsinθ
Figure 2.7 An element of solid angle in spherical coordinates.2
In other words, l = r defines one radian. Since there are 2π radians in a circle, the
conversion between degrees and radians is 1 rad = (360/2π) = (180/π) degrees.
The steradian (sr) is defined in an analogous manner:
The steradian is the solid angle that, having its vertex in the center of a
sphere, cuts off an area of the surface of the sphere equal to that of a
square with sides of length equal to the radius of the sphere.1
Dividing the entire surface area of a sphere by the square of its radius, we
find that there are 4π sr of solid angle in a sphere, and 2π sr in a hemisphere.
Figure 2.7 depicts the solid angle in terms of the planar angle θ and the
rotational (azimuthal) angle φ, where r is sphere radius. The small element of
area dAsph lies on the surface of the sphere. The element of solid angle subtended
by dAsph is expressed as:
dAsph
dω = = sin θd θd φ . (2.4)
r2
ω = sin θd θd φ . (2.5)
φ θ
Θ1/2
Figure 2.8 Right circular cone, with Θ1/2 the cone half angle.
In the most general case, a solid angle can subtend a surface of any shape.
However, in optical systems, which typically have circular apertures, the
envelope of the solid angle is a right circular cone, as shown in Fig. 2.8.
In Fig. 2.8, Θ1/2 is the cone half angle. It is the plane angle between the
centerline of the cone and anywhere on the edge of the cone. It is related to the
solid angle of the cone ω by:
2π Θ1/ 2
ω= dφ
0
sin θd θ ,
0
(2.6)
which results in
A feel for the magnitudes of various solid angles can be obtained by inspecting
Table 2.3. Table 2.4 facilitates conversion from steradians to other units.
Table 2.3 Some objects and corresponding solid angles.
1 sr = 1 rad2
1 sr = 3282.8 deg2
1 sr = 1.1818 × 107 arcmin2
1 sr = 4.2545 × 1010 arcsec2
Both plane angles and solid angles are dimensionless quantities, and their use
can lead to confusion when attempting dimensional analysis. For example, the
simple inverse square law of irradiance (to be discussed in detail in Sec. 2.3.1),
E = I/d 2, appears dimensionally inconsistent. The left side has units W/m2, while
the right side has W/m2sr. It has been suggested that this equation be written E =
I Ωo/d2, where Ωo is the unit solid angle, 1 sr. Inclusion of the term Ωo will render
the equation dimensionally correct, but Ωo will far too often be considered a free
variable rather than a constant equal to 1, which leads to erroneous results.
Current practice suggests that another type of solid angle, the projected (or
weighted) solid angle, is more useful. The symbol for a projected solid angle is
Ω, and the units are also steradians. It is defined as the solid angle ω projected
onto the plane of the observer, as shown by the defining equation:
As before, the example most relevant to optical systems is the right circular
cone depicted in Fig. 2.8, for which the integral may be expressed as
2π Θ1/ 2
Ω = d φ sin θ cos θ d θ , (2.11)
0 0
resulting in
where Θ1/2 is the cone half angle, as before. The plane area subtended by the cone
is Ωr2, and the spherical surface area subtended is ωr2. A couple of special cases
are worthy of mention. For a hemisphere, Ω = π sr, while an entire sphere
subtends 2π sr of a projected solid angle.
In addition to the dimensional concern raised above, there is another good
reason to employ two definitions of a solid angle. For many radiometric
problems, the emitter or receiver is flat, and the projected solid angle Ω is the
proper choice as it requires the inclusion of the cosθ term relating to the
projected area of the surface. In other cases, the emitter or receiver approximates
a point, emitting uniformly in all directions or responding to incoming radiation
equally from all directions. The solid angle ω is appropriate under these
conditions.
For a right circular cone at a half angle Θ1/2 of 90 deg, the projected solid
angle is π according to Eq. (2.12), and ω is 2π according to Eq. (2.7). At the
other extreme, when Θ1/2 is equal to 0 deg, both ω and Ω are zero. For small
angles, the solid angle and projected solid angle differ only by a cosine, are
nearly identical, and are therefore interchangeable in numeric value, if not in
concept. The error incurred by using ω rather than Ω is given in Table 2.5 (note
that ω > Ω for angles greater than 0 deg).
The definitions and symbols presented here have not been universally
applied in the past. One must be very cautious when reading the literature, as
different investigators use the terms and symbols solid angle ω and projected
solid angle Ω interchangeably, incurring predictable confusion and potentially
incorrect results.
Table 2.5 Percentage error when not using a projected solid angle.
f
f #= . (2.12)
D
Good optical systems fulfill the Abbe sine condition with a spherical
wavefront converging to the focal point, and the preferred definition of f/# is
1
f #= . (2.13)
2sin Θ1/2
The numerical aperture (NA) of a system is defined as the sine of the vertex
angle (half angle) of the largest cone of meridional rays that can enter or leave an
optical system or element, multiplied by the refractive index of the medium in
which the vertex of the cone is located. It is generally measured with respect to
an object or image point, and will vary as that point is moved. The defining
equation is
Both f/# and numerical aperture can be related to projected solid angle:
*
For an informative discussion of f/#, see D. Goodman, “The f-word in optics,” Optics &
Photonics News, p. 38, April (1993).
π
Ω= , (2.15)
4( f #) 2
and
π(NA) 2
Ω= , (2.16)
n2
n
f /# = . (2.17)
2NA
T = AΩ . (2.18)
Figure 2.10 shows the interaction between two differential area elements dA1
and dA2 a distance d apart. The area dA1 subtends a solid angle ω1, having apex at
dA2, while dA2 subtends a solid angle ω2, whose apex is at dA1. The surface
normals are shown in the figure by boldface n.
The throughput relationships become
d
n1 n2
Figure 2.10 Area- and solid-angle relationships used to define throughput. (Adapted from
a figure courtesy of William L. Wolfe.)
Figure 2.11 Invariance of throughput for a case in which the source image fills the
detector.
Since these pairings are equal, any of the above pairs can be chosen for
calculation purposes. If the image of the source does not exactly fill the detector
area, care must be taken to determine the proper ΑΩ product to use. The author’s
(Palmer) personal preference is the most often-used pair AoΩdo, inasmuch as the
entrance aperture size Ao and the field of view of the system Ωdo are determinable
characteristics of a radiometer. The next most often used pair is AdΩod, as the
detector size and the f/# of the optics are also measurable characteristics of the
radiometer.
Basic throughput is the name given to the quantity conserved across a
lossless boundary between two media having different indices of refraction. It
can be written as
n 2T = n 2 AΩ , (2.21)
where the subscripts denote the respective quantities in media 1 and 2. Like
throughput, basic throughput is invariant. Most optical systems have both the
object and the image located in the same medium, typically air, so basic
throughput is not often used.
Figure 2.12 depicts the appropriate and inappropriate area- and solid-angle
pairings used to define throughput. The correct area–solid angle pair is shown in
Fig. 2.12(a), the incorrect angle pair is shown in Fig. 2.12(b). Because the
definition of throughput includes two (projected) areas and the distance between
them, a correct pairing has the apex of its solid angle located at the correct area.
The incorrect pairing uses one area twice and ignores the other. The maxim “no
ice cream cones” should be applied.
Let’s look at some examples of the AΩ product. First, a spectrometer: Fig.
2.13 shows the area- and solid-angle product of the entrance slit (typically 1 × 10
mm) and the projected solid angle the collimating lens subtends at the slit, Ωls.
The AΩ product of a spectrometer is usually very small, and narrow spectral
bandwidths are typically employed. Therefore, it is difficult to get much light
through a spectrometer.
A different example of throughput may be found in the common camera. It is
related to the f/# of the lens and the size of the film. In this case, the detector
(film) size is not the overall dimension of the image, but the size of an individual
film grain. The smaller the f/#, the “faster” the camera and the greater the
throughput. Similarly, “fast” film has a larger grain area, permitting a higher
throughput and a shorter exposure time than “slow” film with a smaller grain
area.
(a) (b)
Figure 2.12 Right and wrong area–solid angle combinations for throughput determination.
Slit, Lens
Area A
Ωls
Figure 2.13 Example of the correct area (A) and solid angle (Ωls) product used to
determine throughput in a spectrometer. (Adapted from a figure courtesy of William L.
Wolfe.)
Image
Marginal Ray
Object
Chief Ray
Figure 2.14 Chief and marginal rays in an optical system, shown schematically.
†
Several excellent texts exist; a recent one is E. L. Dereniak and T. D. Dereniak, Geometrical and
Trigonometric Optics, Cambridge University Press, Cambridge (2008).
NP
CHIEF RAY XP
MARGINAL RAY
AS FS
Figure 2.15 Optical system stops and pupils.
The marginal (rim) ray in an optical system is the ray from the object that
originates at the intersection of the object and the optical axis and passes through
the edge of the entrance pupil. It touches the edge (rim) of the aperture stop,
proceeds through the center of the field stop, and the edge of the exit pupil. The
marginal ray intersects the optical axis at the center of the image, defining the
location of the image and the longitudinal magnification. There may be several
intermediate image planes in a complex optical system.
The entrance pupil is the image of the aperture stop in object space (as seen
from the object), while the exit pupil is the image of the aperture stop in image
space (as seen from the image).
Of particular importance to radiometry are the aperture and field stops. The
aperture stop determines how much light may enter the system, while the field
stop determines the system’s angular field of view. In a simple system consisting
of a lens and detector at the rear focal point, the lens is the aperture stop, and
both the entrance and exit pupils are located at the lens, with the same size as the
lens. In more complicated systems, the stops may be internal and separated from
the pupils, as shown in Fig. 2.15.
2.2.1 Radiance
The study of radiometry begins with fundamental units. Radiant energy has the
symbol Q and has as its unit the joule (J). Radiant power, also known as radiant
flux, is energy per unit time (dQ/dt), has the symbol Φ, and is measured in watts
(W). These definitions give no indication of the spatial distribution of power in
terms of area or direction. Radiance is the elemental quantity of radiometry,
power per unit area, and per unit projected solid angle. It is a directional quantity;
it can come from many points on a surface that is either real or virtual; and
because it is a field quantity, it can exist anywhere. The symbol for radiance is L
and the units are W/m2sr.
The defining equations are
ΔΦ d 2Φ d 2Φ d 2Φ
L = lim = = = , (2.23)
ΔAs , Δω→ 0 ΔA cos θΔω
s dAs cos θd ω dAs d Ω dAp d ω
where θ is the angle between the normal to the source element and the direction
of observation as shown in Fig. 2.16.
Radiance is also associated with a source, either active (thermal or
luminescent) or passive (reflective), as discussed further in Chapter 4. Because
radiance may be evaluated at any point along a beam, it is associated with
specific locations within an optical system, including image planes and pupils.
Other radiometric units may be derived from radiance by integrating over
area and/or solid angle. Integration over solid angle yields irradiance (arriving at
a location, such as a sensor) or radiant exitance (leaving a location, such as a
source), both of which are expressed in W/m2. Integration over area yields
radiant intensity expressed in W/sr. Integration over both area and solid angle
yields radiant power in watts.
If rays are traced across a lossless boundary between two materials having
different indices of refraction as shown in Fig. 2.3, the solid angle changes
according to Snell’s law. Taking this change into account, the quantity Ln–2 is
seen to be invariant across the boundary. This quantity is called basic radiance. It
is useful for calculations when an object and its image are located in spaces with
different indices of refraction.
In the absence of sources or sinks along the path of a beam, power along a
beam is conserved. Since it was previously demonstrated that throughput is
conserved in an optical system, the radiance must also be invariant in order for
conservation of power (energy per unit time) to be obeyed. The results of this
invariance of radiance are significant:
(1) The radiance of the image at the detector plane of a camera (film or array
device) is the same as the radiance of the scene if there are no
transmission losses due to atmosphere and optics; and
(2) The radiance at the focal plane of a radiometer (imaging or point) is the
same as that of the target, if there are no transmission losses due to
atmosphere and optics.
Note that the transmission of atmosphere and optics is not likely to be unity
(perfect transmission); however, results (1) and (2) greatly simplify radiometric
calculations.
dω
θ
n
Figure 2.16 Radiance from area element dA, tilted at angle θ from surface normal, n.
The defining equation for radiance can be inverted and integrated over area
(and in the most general case, a projected solid angle) to determine the power in
an optical system:
ΔΦ d Φ
M = lim = . (2.25)
ΔAs → 0 ΔA
s dAs
M = L dΩ . (2.26)
π
Φ = M dAs . (2.27)
A
Figure 2.17 further illustrates the concept of radiant exitance. The relationship
between radiant exitance and radiance is complex, depending on the angular
distribution of radiance L(θ,φ). This is illustrated in Fig. 2.18 and described
mathematically in the following equations:
where r is the radius of the hemisphere, as shown in Fig. 2.18, dA1 is an element
of the emitting area, dA2 an element of the receiving area, and θ is the angle
between the surface normal of dA1 and the direction of propagation.
Making a substitution for dA2 using polar coordinates,
dA2 = r 2 sin θd θd φ
results in
d 2 Φ = L(θ, φ)dA1 sin θ cos θd θd φ . (2.29)
2π π /2
Φ = dφ L ( θ, φ ) A1 sin θ cos θd θ . (2.30)
0 0
π2
sin 2 θ
Φ = 2πLA1 . (2.31)
2 0
M = πL . (2.32)
r2sinθdθdφ=dA2
r
θ
dA1
Figure 2.18 Relationships between emitting and receiving areas used in deriving the
relationship between radiance and radiant exitance.
2.2.3 Irradiance
Irradiance (radiant incidence) is the opposite of radiant exitance, in that it is the
power per unit area that is incident on a surface. Its symbol is E and its units are
again W/m2:
ΔΦ d Φ
E = lim = . (2.33)
ΔAs → 0 ΔA
s dAs
E = Ld Ω . (2.34)
π
Φ = EdA . (2.35)
A
ΔΦ d Φ
I = lim = . (2.36)
Δωs → 0 Δω d ωs
s
I = LdA , (2.37)
A
and radiant power can be derived from intensity by integrating over solid angle
ω. (Note that the definition of intensity does not include an area term; therefore,
the integration is performed over solid angle, without a cosine projection.)
Φ = Id ω . (2.38)
ω
In the case of the isotropic point source, the radiant power is 4π times the
intensity. Conversely, intensity may be found by dividing the total emitted power
Φ by 4π.
A word of warning: intensity is the most problematic radiometric quantity,
because it means different things to different people in different but related
fields. Laser physicists are prone to use it as the square of the electric field
strength, with units W/m2. Atmospheric scientists and heat transfer engineers use
the term “specific intensity,” frequently omitting the “specific,” to mean
W/m2sr. Some scientists and engineers equate intensity simply to power, and a
few use it to describe spectroscopic line strengths. Which usage is correct?
Φ I
E= = . (2.39)
A d2
The relationship expressed by Eq. (2.39) presumes that the area shown in
Fig. 2.21 lies normal to the optical axis. If this is not the case and the surface is
tilted at an angle θ to the optical axis, a cosine θ factor must be included in the
equation:
I cos θ
E= . (2.40)
d2
I A
d θ
n
Figure 2.22 Area A tilted at angle θ from the axis.
I
Ex = .
D2
At position y, the distance from the source has increased such that d=D/cosθ.
If the target at y is perpendicular to the vector between source and target, the
irradiance at y is:
I I cos 2 θ
Ey = = .
d2 D2
Is
θ
d D
y x
If the target is now rotated so that it is parallel to surface x-y (the floor), an
additional cosθ term is introduced, resulting in
I cos3 θ
Ey = ,
D2
or more generally,
I cos3 θ
E= , (2.41)
D2
which is the cosine3 law. Note that two cosines were introduced to account for
the increased distance from the source; the third accrued from the projected area
of the target.
L (θ,φ) = constant.
the power received from a Lambertian radiator is proportional only to cosθ, the
angle of observation, and by extension, the projected area of the source.
Ls
θ
d
D
y x
Recall that the projection of the target onto the floor added the third cosθ
term; the projection of the source area in the direction of point y adds the fourth.
For a Lambertian source, the relationship between intensity (point source)
and radiance (extended source) is I = LAproj. Substituting,
LAp cos 4 θ
E= , (2.42)
D2
Θ 1/2
An expression for the irradiance at the sensor for this particular geometry has
been derived by Foote (1915).9,10 Foote’s formula is shown below, and the
corresponding plot using f/# as a parameter is shown in Fig. 2.26. For this
particular geometry, irradiance is least likely to follow the cosine4 law in lower
f/# systems. Note that at normal incidence (θ = 0), the equation reduces to
E = πLsin2θ.
1 + tan 2 θ − tan 2 Θ1 2
πL
E= 1 − 1/ 2
.
2 tan θ + 2 tan θ(1 − tan Θ1 2 ) + sec Θ1 2
4 2 2 4
The physical situation described by the equation is shown in Fig. 2.27, where
dA1 and dA2 are the differential area elements, d is the distance between them,
and θ1 and θ2 are the angles between the normals to the area elements and the
optical axis.
As the equation indicates, the radiant power received at surface 2 depends
directly on the radiance L(θ,φ), the area elements, the angles between area
normals and the optical axis, and inversely on the square of the distance between
the surfaces. If the distance squared is much larger than the largest area, i.e.,
d 2 >> (A1 or A2), the differential areas may be replaced with the actual areas, and
Eq. (2.44) can be applied to calculate power:
Because d 2 is much greater than the size of any area element, the variations
in θ and φ are small, and the term L(θ,φ) need be evaluated at only one particular
set of angles. If either area is appreciable in relation to d 2, Eq. (2.44) cannot be
used.
The integral form of the equation of radiative transfer looks much more
formidable:
θ1 d
θ2
dA2
dA1
To utilize this integral form, the following factors must be taken into
account:
(1) The angles θ1 and θ2 may vary from one part of area A1 or area A2 to
another.
(2) The distance d may also vary from one part of area A1 or area A2 to
another.
(3) The angular variation in radiance L(θ,φ) may be significant.
cos θ1 cos θ2
Φ1− 2 = L dA1dA2 . (2.46)
A2 A1
d2
These simple equations are the logical starting points for all radiometric
engineering calculations, as they provide first-cut, back-of-the-envelope answers.
In many instances, they are all we need. In order to fully understand a particular
application, assumptions must be tested and the errors incurred by their use
assessed.
The Lambertian approximation is relatively simple to verify if we possess
hard data about the directionality of the source. The on-axis assumption is also
easy to verify. As discussed in Sec. 2.3, the distance between the two area
elements must be at least 10× the maximum linear dimension of the largest
element in order for the inverse square law to be good to 1%. If the distance
between the two is increased to 20× the maximum dimension, the uncertainty in
applying the law reduces to 0.1% or less.
Φ1→ 2
F= , (2.49)
Φ1
where Φ1 is the power leaving surface 1 and Φ1→ 2 is the power reaching surface
2 from surface 1. Both power terms are dimensionless. The radiant power terms
are further defined as
Φ1 = M 1 A1 ,
where M1 is the radiant exitance in W/m2 leaving surface 1, and A1 its area; and
M 1 cos θ1 cos θ2
π
Φ1→ 2 = dA1dA2 ,
d2
where the radiance term outside the integral is obtained from Eq. (2.32), itself
dependent upon the Lambertian approximation. The fraction of radiant power
leaving surface 1 that arrives at surface 2 is
1 cos θ1 cos θ2
F12 =
πA1 d2
dA1dA2 ,
which is the configuration factor. The power transferred from surface 1 to surface
2 then becomes
Φ1→ 2 = Φ1 F12
= M 1 A1 F12
= πL1 A1 F12 .
and
E2 = L1Ω12 . (2.51)
1 r
2
Fd 1−2 = Fd 1− 2 =
(h r )
2
+1 h
A2 A2
r
r
h
h
dA1
dA1
(a) (b)
Figure 2.28 Configuration factor examples: (a) Planar element parallel to circular disk,
and (b) planar element to sphere. (Adapted from Ref. 11.)
Is
Ed = ,
d2
where d is the distance between source and detector. Without a lens, [Fig.
2.29(a)], assuming no transmission losses in the intervening medium, the power
at the detector is simply the irradiance multiplied by the available sensitive area:
Φ d = Ed A2 . (2.52)
I s A2
Φd = , (2.53)
( D + S )2
(a) (b)
Φ d = Ed A1 , (2.54)
τlens I s A1
Φd = . (2.56)
D2
Note that in this case, the aperture that limits the flux into the system is A1,
the aperture stop, and that A2 is the field stop, limiting the detector’s field of
view. The size of A2 is unimportant as long as it does not vignette the source’s
image at the detector.
The difference in received power between the two expressions in terms of
irradiance is expressed as
τlens A1
G= , (2.57)
A2
where G is the gain of power on the detector. To maximize G, make the lens
transmission and the area ratio as large as possible, while not vignetting the
source image.
To determine the effect of a lens on the same instrument configuration with
an extended source, begin with Eq. (2.47):
Φ d = Ls AΩ .
At area A = A2, the appropriate solid angle is subtended by area A1. This solid
angle is expressed as in Eq. (2.12) by
where Θ1/2 is the cone half angle. Assuming that it is small, the approximation
A1/S2 may be used for Ω12, so that
A1 A2
AΩ = , (2.58)
S2
and
Ls A1 A2
Φd = . (2.59)
S2
The same approach can be pursued with the other area–solid angle
combination, that is, with A1 and Ω21. In that case, the solid angle is
approximately A2/S2, and Φd is again obtained by Eq. (2.59). Inserting a lens at A1
limits the power by the transmission of the lens, so that
τlens Ls A1 A2
Φd = . (2.60)
S2
π
Ω sd = π sin 2 Θ1/2 = , (2.61)
4( f #) 2
where f/# is defined in Eq. (2.13). Considering the geometry in the figure,
a2
sin 2 Θ1/ 2 = .
(a 2 + b 2 )
a2 πL
Ed = πL sin 2 Θ1/ 2 = LΩ sd = πL = = πL( NA) 2 . (2.62)
(a + b ) 4( f #) 2
2 2
a Θ1/2
Ad
b
Ls
Figure 2.30 On-axis Lambertian disc, irradiance measured at detector of area Ad.
If the distance b is far greater than 2a, the linear dimension of the source,
then the inverse square law holds and Ed may be approximated as πLa2b–2. The
error incurred using a2b–2 rather than a2(a2 + b2)–1 is less than 1% when the
diameter-to-distance ratio is less than 0.1. Under these conditions, source
intensity Is may be substituted for πLa2 (the radiance times the area of the source)
so that
Is
I s = LAs Ed = .
b2
Table 2.8 Irradiance at detector as a function of source distance for a Lambertian disc.
a
Θ1/2
Ad
Ls
integrating Eq. (2.26) over a hemisphere, and the irradiance at the detector
calculated as a function of the half angle. To illustrate, take the relatively simple
example in which Ls = Locosθ. In this case, the radiant exitance is
M s = Lo cos θd Ω , (2.63)
2π π /2
M s = Lo d φ sin θ cos 2 θd θ , (2.64)
0 0
which results in
2πLo
Ms = . (2.65)
3
2πLo
Ed =
3
(1 − cos3 Θ1/ 2 ) . (2.66)
a2 a2
sin 2 Θ1/ 2 = = , (2.67)
(a + b) 2 a 2 + 2ab + b 2
Table 2.9 Irradiance at the detector as a function of source distance for a Lambertian
sphere, measured from a surface.
a2
Ed = πL . (2.68)
a 2 + 2ab + b 2
Table 2.9 summarizes irradiance at the detector for a variety of cases. Note
that when b >> 2a, the inverse square law applies and the irradiance from the
sphere is the same as that from the disc, above.
If the source–detector distance is measured from the center of the sphere, as
shown in Fig. 2.32, the sine of the half angle is always a/b. The irradiance at the
detector is therefore
a 2 LAs I s
Ed = πL = 2 = 2. (2.69)
b2 b b
Equation (2.69) reveals an interesting result: the inverse square law holds for any
sphere and at any source–detector distance, as long as the surface is Lambertian
and the distance is measured from the center of the sphere. This counterintuitive
result simplifies calculation; results are shown in Table 2.10.
a
Θ1/2
Ad
Ls
Figure 2.32 On-axis Lambertian sphere, irradiance measured from the center.
Table 2.10 Irradiance at the detector as a function of source distance for a Lambertian
sphere, measured from the center.
d Φ LdA
dE = = . (2.70)
dA 4 R 2
dA1
d
θ1
R θ2 dA2
R
Figure 2.33 The integrating sphere. (Adapted from Ref. 15 with permission from John
Wiley & Sons, Inc.)
This result means that irradiance within the sphere, for any area element dA,
is independent of position θ within the sphere and is dependent only on sphere
radius and radiance L. In other words, irradiance is constant over the sphere. This
fact makes the integrating sphere useful as a uniform radiance source.
If a source with power Φ is placed into the sphere (through a “port” in the
sphere), the radiance of the sphere wall L (assumed to be Lambertian) can be
determined as
Eρ
L= , (2.71)
π
where ρ is a property of the sphere coating material called its reflectance (to be
discussed in detail in Chapter 3). Combining Eqs. (2.70) and (2.71) and solving
for dE, we obtain
EρdA
dE = . (2.72)
4πR 2
ρΦ
E= . (2.73)
4πR 2 (1 − ρ)
ρΦ
L= , (2.74)
πAsph [1 − ρ(1 − f ) ]
in which f is the ratio of the total port area to that of the sphere. (A sphere may
have several ports.) Thus, real spheres are not particularly efficient unless
reflectance is high and the total port area is kept small. Table 2.11 details some of
the many uses of integrating spheres.
Φ m 4 × 10−6
Em = = = 5 × 10−6 W/m 2 .
Am πD 2
The inverse square law applies due to the source distance, and as the source
is on axis, no cosine term is required. Inverting Eq. (2.39) to calculate intensity,
we have
I
DET
F/2,
f/2,
12
f =f=1m
1
d= 1012
d=10
Figure 2.34 Hypothetical distant star and system used to measure irradiance.
σT 4
L= , (2.75)
π
and its value is 2 × 107 W/m2sr. Consider the geometry in Fig. 2.35, where Ap is
the projected area of the sun. The sun’s diameter is 1.4 × 109 m, so according to
Table 2.2, its projected area is (πdsun2)/4, or 1.54 × 1018 m2. At earth-sun distance
d = 1.5 × 1011 m, the solid angle subtended by the sun at the detector is Ωsd = 6.8
× 10–5 sr. Noting also that the sun subtends approximately 32 minutes of arc
(arcmin), Ωsd may also be calculated as πsin2(16 arcmin), which produces the
same result. The irradiance at the detector, area A in the figure (assumed to be
placed at the top of the atmosphere, therefore no atmospheric transmission loss),
is
Note that the diameter-to-distance ratio is substantially less than 0.1, and the
inverse square law may be applied. Calculating intensity as in the example
above:
Ap
A
Irradiance may also be obtained in another way. Consider that the power
delivered to the detector with area A may be represented by
LAp A
Φ d = LAp Ω ds = , (2.78)
d2
where Ωds is the solid angle subtended by the detector, and that
Φ d LAp
Ed = = 2 = 1368 W/m 2 . (2.79)
A d
Applying the inverse square law as in Eq. (2.77), we obtain Isun = 3.08 × 1025
W/sr. Note that though the numbers are not identical, they are very close. The
value of 1368 W/m2 is referred to as the solar constant, and is specifically defined
as the irradiance falling upon a 1 m2 unit surface (hypothetical surface) at the
mean earth–sun distance. The solar constant has wide application in fields
including remote sensing. It is often given the symbol Eo.
Note that the total intensity of the sun has to do with the power radiated into
4π sr, the solid angle of a sphere as referenced in Table 2.3. Table 2.12 provides
relevant calculations related to solar power and intensity.
2.9.2.1 Calculation
This concept is analogous to that of the solar constant, whose 1368 W/m2 are
incident upon the moon as well. If the moon is assumed to be Lambertian, with a
reflectance of 0.2, its radiance is
Eo ρ
Lmoon = = 87 W/m 2sr . (2.80)
π
At the earth’s surface, the angular subtenses of the moon and the sun are the
same, approximately 32 arcmin. This means that ΩME, the solid angle subtended
by the moon at the earth, is equal to Ωsd, above, with a value of 6.8 × 10–5 sr.
Neglecting the relatively minimal distance between the top of earth’s atmosphere
and its surface, the irradiance produced by the moon at the top of earth’s
atmosphere is
Φ M →E
Emoon , TOA = = Lmoon Ω ME = 5.9 × 10−3 W/m 2 . (2.81)
ApE
Note that in the above equation, ApE is the projected area of the earth,
analogous to the projected area of the sun discussed earlier.
The numbers are the same because the solid angles subtended are the same.
Assuming an atmospheric transmission of 0.75, the solar irradiance at the
earth’s surface is this factor multiplied by the solar constant
Eearth ρ
Learth = = 65 W/m 2sr .
π
This interesting result means that the radiance of an “average” sunlit scene‡ is
the same as the apparent radiance of the moon. It also means that in photography,
the same exposures can be used to photograph both. Exposure parameters should
be set during the daytime and applied to night photography. If the moon is
photographed through a telescope, exposures should be increased to compensate
for transmission losses within the instrument. In addition, given the factor of
2.3 × 105 difference between solar and lunar irradiances, photographing a
moonlit scene requires significantly longer exposures than are needed for
daylight illumination.
A point about assumptions should be made, specifically, that the moon is not
a strict Lambertian surface. It is somewhat retroreflective, as though covered
with Scotchlite.™ Simple measurements made by Palmer with a silicon detector
indicate that the apparent intensity of the full moon is more than ten times that of
the quarter moon. When viewed with a telescope or binoculars, the edge appears
a bit brighter than the rest. The lunar surface is dusted with small glassy
spheroids, ejecta from meteorite collisions. Its reflectance is approximately 0.08,
somewhat less at shorter wavelengths and somewhat more at longer wavelengths.
Φ s →c
Ec = = Ls Ω sc ,
Ac
where Ωsc is the solid angle the sun subtends at the collector, 6.8 × 10–5 sr.
Choosing a target diameter (or linear dimension) and system focal length f so that
Ωsc = Ωtc, we have
At
Ωtc = and At = Ωtc f 2 .
f2
‡
Eastman Kodak has shown through extensive research that the reflectance of an average scene is
18%; all exposure meters are calibrated using this assumption (J. M. Palmer, 2005).
Ac Ωtc
Ls = 2 x 107 Wm–2sr–1
At
Ωsc=Ωtc
f
Figure 2.36 The “solar furnace.”
Φ c →t Ec Ac E A
Et = = = c c2 .
At At Ωtc f
Therefore, the target irradiance is the product of the source radiance and solid
angle the collector subtends at the target. That solid angle may also be
characterized [Eq. (2.15)] as
π
Ω ct = ,
4( f #) 2
so that
πLs
Et = , (2.82)
4( f #) 2
As Ac At
m
f # ′ = f # 1 + , (2.83)
m
p
where magnification m is the ratio of image height to object height, and has
values between 0 and infinity. The term mp is pupil magnification, the ratio of the
diameter of the exit pupil to the diameter of the entrance pupil, and has values
between 0.5 and 2. For a single lens or mirror, it is always 1. Substituting f/#′ for
f/# in Eq. (2.82) we obtain
πL
Et = 2
. (2.84)
m
4 ( f # ) 1 +
2
mp
πL
Et = . (2.85)
4 ( f # ) (1 + m )
2 2
The expressions for irradiance at the target show us that image irradiance
decreases as the in-focus object is moved closer to the camera. In order to
maintain focus, the detector must be moved backward, which decreases the solid
angle of the lens as seen from the detector Ωct by a factor of four.
Φ sky →earth 2π π/ 2
Eearth = = Lsky Ω sky −earth = Lsky d φ sin θ cos θd θ
Aearth 0 0
2
Eearth = πLsky sin (90 deg)
Eearth = 157 W/m 2 .
This is a factor of 6 or 6.5 less than the irradiance received from the sun on a
clear day (1000 W/m2), which explains why flat-plate solar collectors continue to
function well on a cloudy day (provided that the clouds are “conservative”
scatterers.) By comparison, on a clear day, the diffuse solar irradiance (excluding
the direct beam) can be as high as 50 to 100 W/m2 due to scattering.
(1) If the image of the extended source overfills the field of view of the
detector with area Ad, the distance d is unimportant;
(2) If the source is Lambertian, the angle between source and optical axis is
unimportant; and
(3) If the detector or radiometer with area Ad is not placed exactly at the
focal distance f, it doesn’t matter.
Ao πDo 2
Φ d = Ls Ad Ωod = Ls Ad 2
= Ls Ad ,
f 4f2
which equates to
πLs Ad
Φd = .
4( f #) 2
Ls
θ
Ad
d f
Figure 2.38 Near extended source.
The equation for illuminance on the screen resembles the camera equation,
Eq. (2.85), with the addition of a cos4θ term to account for the off-axis angle to
the screen as seen from the projector:
τo πLv cos 4 θ
Ev = , (2.86)
4 ( f # ) (1 + m )
2 2
τo πf v (1 − A2 ) L(θ, φ) cos n θ
Et = . (2.87)
4 ( f # ) (1 + m )
2 2
Note that this expression does not take into account the spectral nature of the
radiation, to be discussed in greater detail in Chapter 4, nor does it account for
the transmission of the atmosphere.
τo f v (1 − A2 ) I cos n θ
Et = , (2.88)
d2
Φ = LAΩ ,
and
E = LΩ .
Further, the Lambertian approximation is okay for most emitters, but angles
must be considered when applying it to metals. It is alright for matte reflectors,
and no good at all for specular reflectors.
The choice of solid angle for radiometric calculations is an important one.
When the source or receiver is isotropic, solid angle ω may be used. When the
source or receiver is Lambertian, projected solid angle Ω is the correct choice.
A final comment on sources deserves mention. These “notes” (lyrics by Jon
Geist and Ed Zalewski, NIST, ca. 1982) are to be sung to the theme song from
the ancient television series, “Mr. Ed:”
References
1. Guide for the Use of the International System of Units (SI), NIST SP811,
U.S. Government Printing Office (1995).
2. F. E. Nicodemus, Self-Study Manual on Optical Radiation Measurements:
Part I–Concepts, NBS Technical Note 910-1, p. 68, NBS, U.S. Government
Printing Office, Washington, D.C. (1976).
3. C. Wyatt, Radiometric Calibration: Theory and Methods, Academic Press,
New York (1978).
4. S. Chandrasekhar, Radiative Transfer, Clarendon Press, Oxford (1950),
reprinted Dover (1960).
5. P. Drude, Lerhbuch der Optik (1900), translated into English and reprinted as
The Theory of Optics, Longmans, Green, New York (1902).
6. B. N. Taylor, The International System of Units (SI), NIST SP330, U.S.
Government Printing Office, Washington D.C. (1991).
7. J. M. Palmer, “Getting intense on intensity,” in Metrologia 30(4), pp. 371–
372 (1993).
8. J. M. Palmer, “Intensity,” in Optics & Photonics News, p. 6, February
(1995).
9. P. D. Foote, “Illumination from a radiating disc,” Bulletin of the Bureau of
Standards, NBS, 12, p. 583 (1915).
10. R. Kingslake, Optical System Design, Academic Press, New York (1983).
11. B. T. Chung and P. S. Sumitra, “Radiation shape factors from plane point
sources,” J. of Heat Transfer 94(3), pp. 328–330 (1972).
12. E. M. Sparrow and R. D. Cess, Radiation Heat Transfer, Brooks/Cole,
Belmont, California (1970).
13. P. Moon, The Scientific Basis of Illuminating Engineering, McGraw-Hill,
Dover, New York (1936).
RETROREFLECTION
DIFFUSE
INCIDENT TRANSMISSION
BEAM
INCIDENT
BEAM
DIFFUSE
REFLECTION
61
A continuing dialog over terminology has taken place, particularly over the
suffixes -ance and –ivity. 1,2,3,4,5 The usage here reserves terms ending with -ivity
(such as transmissivity, absorptivity, and reflectivity) for properties of a pure
material, while the suffix –ance is used when the characteristics of a specific
sample are described. One can then distinguish between the reflectivity of pure
aluminum (as calculated from the complex index of refraction n and κ) and the
reflectance of a particular specimen of 6061 aluminum with surface structure
associated with rolling or machining and with a natural oxide layer.
The adjective spectral refers to a characteristic at a particular wavelength and
is indicated as a function of wavelength λ, i.e., τ(λ), ρ(λ) or α(λ). For example,
spectral transmittance τ(λ) is often plotted against wavelength λ for a colored
filter. The absence of “spectral” implies integration over all wavelengths,
weighted by a source function.
3.2 Transmission
Transmission is the process by which incident radiant flux leaves a surface or
medium from a side other than the incident side (usually the opposite side). The
spectral transmittance τ(λ) of a medium is the ratio of the transmitted spectral
flux Φλt to the incident spectral flux Φλi:
Φ λt
τ(λ) = . (3.1)
Φ λi
Total transmittance τ is the ratio of the total transmitted flux Φt to the total
incident flux Φi:
∞
Φ τ ( λ ) Φ λ i dλ
τ= t = ≠ τ(λ) dλ .
0
∞
(3.2)
Φi
λ
Φ λ i dλ
0
Note particularly that the total transmittance is not the integral over
wavelength of the spectral transmittance; it must be weighted by the incident
source function Φλi.
The transmittance may also be described in terms of radiance as follows:
τ=
0 2 π sr
Ltλ dΩt dλ
, (3.3)
∞
0 2 π sr
Liλ dΩi dλ
where Lλi is the spectral radiance Lλi(λ;θi,φi) incident from direction (θi,φi), Lλt is
the spectral radiance Lλt(λ;θt,φt) transmitted in direction (θt,φt), and dΩ is the
elemental projected solid angle sinθ cosθdθdφ.
Geometrically, transmittance can be classified as specular, diffuse, or total,
depending upon whether the specular (regular) direction, all directions other than
the specular, or all directions are considered. The bidirectional transmittance
distribution function (BTDF, symbol ft and units sr–1) relates the transmitted
radiance to the radiant incidence (irradiance) as
dLt ( θt , φt ) dLt ( θt , φt )
f t ( θi , φi ; θt , φ t ) ≡ = . (3.4)
dEi ( θi , φi ) Li ( θi , φi ) d Ωi
3.3 Reflection
In reflection, a fraction of the radiant flux incident on a surface is returned into
the hemisphere whose base is the surface containing the incident radiation. The
reflection can be specular (in the mirror direction), diffuse (scattered into the
entire hemisphere), or a combination of both. Table 3.1 shows a wide range of
materials that have different goniometric (directional) reflectance characteristics.
Spectral reflectance is defined at a specific wavelength λ as
Φ λr
ρ( λ ) = , (3.5)
Φ λi
Material γ
Scatter σ Structure Example
classification (deg)
Exclusively none 0 ≅0 none Mirror
reflecting
materials micro Matte aluminum
weak ≤0.4 ≤27
macro Retroreflectors
τ=0 none Lacquer & enamel
coatings
micro Paint films, BaSO4,
strong > 0.4 > 27
PTFE
macro Rough tapestries, road
surfaces
Weakly transmitting,
Sunglasses, color filters,
strongly reflecting none 0 ≅0 none
cold mirrors
materials
micro Matte-surface color
weak ≤ 0.4 ≤ 27 filters
τ ≤ 0.35 macro Glossy textiles
none Highly turbid glass
strong > 0.4 > 27 micro Paper
macro Textiles
Strongly none 0 ≅0 none Window glass
transmitting
materials none Plastic film
micro Ground glass
weak ≤ 0.4 ≤ 27
τ > 0.35 macro Ornamental, prismatic
glass
none Opal glass
micro Ground opal glass
strong > 0.4 > 27 macro Translucent acrylic
plastic with patterned
surface
while the total reflectance ρ is the ratio of the reflected flux Φr to the incident
flux Φi:
∞
Φr ρ(λ) Φ λi d λ
ρ= = ≠ ρ( λ ) d λ .
0
∞
(3.6)
Φi
λ
Φ λi d λ
0
As shown in Fig. 3.3, the polar angle θ is measured from the surface normal,
z. The azimuth angle φ is measured from an arbitrary reference in the surface
plane, most often the plane containing the incident beam. The subscripts i and r
refer to the incident and reflected beams, respectively.
dΩr
dΩi
θr
θi
dA y
φi
φr
x
Figure 3.3 Geometrical definitions for BRDF (Adapted from Ref. 7).
Nicodemus et al.7 integrated over various solid angles and applied the earlier
work of Judd8 to obtain nine goniometric reflectances and nine goniometric
reflectance factors. These are listed in Tables 3.2 and 3.3. In these tables, the
term “directional” refers to a differential solid angle dω in the direction specified
by (θ,φ). “Conical” refers to a cone of finite extent centered in direction (θ,φ); the
solid angle Ω of the cone must also be specified. The reflectances are illustrated
in Fig. 3.4.
Bidirectional
R (θi , φi ; θr , φr ) = πf r ( θi , φi ; θr , φr )
reflectance factor
Directional-conical π
R(θi , φi ; ωr ) = f r ( θi , φi , θr , φr )d Ω r
reflectance factor Ωr ωr
Directional-
hemispherical R(θi , φi ;2π) = f r ( θi , φi , θr , φr )d Ω r
2π
reflectance factor
Conical-directional π
R (ωi ; θr , φr ) = f r ( θi , φi , θr , φr )d Ωi
reflectance factor Ωi ωi
Biconical π
R(ωi ; ωr ) = f r ( θi , φi , θr , φr )d Ω r d Ωi
reflectance factor* Ωi ⋅ Ω r ωi ωr
Conical- 1
hemispherical R (ωi ;2π) =
Ωi
ωi 2π
f r ( θi , φi ; θ r , φ r ) d Ω r d Ω i
reflectance factor*
Hemispherical-
directional R (2π; θr , φr ) = f r ( θi φi ; θ r φ r ) d Ω i
2π
reflectance factor
Hemispherical- 1
conical reflectance R (2π; ωr ) =
Ωr
2 π ωr
f r ( θi , φi ; θ r , φ r ) d Ω r d Ω i
factor*
Bihemispherical 1
π 2 π 2 π
R (2π;2π) = f r (θi , φi ; θr , φr )d Ω r d Ωi
reflectance factor
*
Configurations that are measurable in practice.
Bidirectional
dρ(θi , φi ; θr , φr ) = f r ( θi , φi ; θ r , φ r ) d Ω r
reflectance
Directional-conical = f r ( θi , φi ; θ r , φ r ) d Ω r
ρ(θi , φi ; ωr )
reflectance ωr
Directional-
hemispherical ρ(θi , φi ;2π) = f r (θi , φi ; θr , φr )d Ω r
2π
reflectance
Conical-directional d Ωr
Ωi ωi
dρ(ωi ; θr , ϕr ) = f r ( θi , φi ; θ r , φ r ) d Ω i
reflectance
Biconical 1
Ωi ωi ωr
ρ(ωi ; ωr ) = f r ( θi , φi ; θ r , φ r ) d Ω r d Ω i
reflectance*
Conical- 1
hemispherical ρ(ωi ;2π) =
Ωi
ωi 2π
f r ( θi , φi ; θ r , φ r ) d Ω r d Ω i
reflectance *
Hemispherical- d Ωr
directional dρ(2π; θr , φr ) =
π 2π
f r (θi φi ; θr φr )d Ωi
reflectance
Hemispherical- 1
π 2 π ωr
conical ρ(2π; ωr ) = f r (θi , φi ; θr , φr )d Ω r d Ωi
reflectance*
Bihemispherical 1
π 2 π 2 π
ρ(2π;2π) = f r (θi , φi ; θr , φr )d Ω r d Ωi
reflectance
*
Configurations that are measurable in practice.
In both Tables 3.2 and 3.3, configurations containing a directional term are
considered theoretical, as dΩ→0.
3.4 Absorption
Absorption is the process in which a fraction of the incident radiant flux is
converted to another form of energy, usually heat. Absorptance is the fraction of
incident flux that is absorbed. Spectral absorptance is defined at a specific
wavelength λ as
Φλa
α (λ ) = , (3.8)
Φλi
with the subscripts denoting absorbed and incident power, respectively. Total
absorptance is defined as
∞
Φ α ( λ ) Φ λi d λ
α= a = ≠ α (λ ) d λ .
0
∞
(3.9)
Φi
λ
Φ λi d λ
0
τ + ρ + α =1. (3.10)
The above statement assumes integration over all wavelengths and directions. In
the absence of wavelength-shifting effects (such as luminescence or Raman
scattering), this relationship is also valid for any specific wavelength:
2
n′ cos θ − n cos φ
ρp =
n′ cos θ + n cos φ
2
, (3.12)
n cos θ − n′ cos φ
ρs =
n cos θ + n′ cos φ
and
2
2n cos θ n′ cos φ
τp =
n′ cos θ + n cos φ n cos θ
2
, (3.13)
2n cos θ n′ cos φ
τs =
n cos θ + n 'cos φ n cos θ
where the subscripts p and s represent the two polarization states, n and θ are on
the incident side of the interface, and n′ and φ are on the transmitted (or reflected)
side. The total transmittance and reflectance for unpolarized light is the average
of the two polarized components
ρ p + ρs
ρT =
2
and (3.14)
τ p + τs
τT = .
2
Figure 3.5 shows reflection and transmission curves for a single surface of a
nonabsorbing optical material with an index of refraction of 2. The three curves
represent s- and p-polarization states as well as total polarization. To compute the
curves for absorbing media, substitute n ± iκ for n in Eqs. (3.12) and (3.13),
where κ = αλ 4π . Since refractive index n is wavelength dependent, the
calculated reflectance and transmittance are also.
In Fig. 3.6, we see a partially transparent plane slab of an optical material.
Reflection, transmission, and absorption are all present.
The Fresnel equations are greatly simplified at normal incidence, in which
θ = φ = 0. For a nonabsorbing material, the reflectance and transmittance at a
single surface are
2
n′ − n 4nn′
ρss = τss = . (3.15)
( n′ + n )
2
n′ + n
1
SINGLE SURFACE REFLECTANCE 0.9
0.8
0.7
0.6
(a) 0.5
0.4
S
0.3 TOTAL
0.2
P
0.1
0
0 10 20 30 40 50 60 70 80 90
ANGLE FROM NORMAL (deg)
1
SINGLE SURFACE TRANSMITTANCE
0.9
TOTAL P
0.8
0.7
0.6
(b) 0.5 S
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50 60 70 80 90
ANGLE FROM NORMAL (deg)
Φi
Φr Φt
Φa
Φ t (1 − ρss ) e (1 − ρss ) τi .
2
−α′ x 2
τ= = 2 −2 α′ x
= (3.17)
Φi 1 − ρss e 1 − ρss 2 τi2
2n
τ= . (3.18)
n2 + 1
For absorptance:
α= = = . (3.20)
Φi 1 − ρss e −α′ x 1 − ρ ss τi
When the optical thickness is large enough, the material becomes opaque and
the transmittance goes to zero. In this case, the reflectance ρ approximates the
single surface reflectance ρss, and the absorptance α approaches 1 – ρ. For the
opposite case, in which the optical thickness approaches zero, the material
becomes transparent and the following relationships hold:
ρss (1 − ρss )
2
ρ = ρss + , (3.21)
1 − ρss 2
(1 − ρss )
2
τ= . (3.22)
1 − ρss 2
Figure 3.7(a) BRDF of rough aluminum at incidence angle of 33 deg. (Reproduced from
Ref. 9 with permission.)
Figure 3.7(b) BRDF of rough aluminum at incidence angle of 33 deg. (Reproduced from
Ref. 9 with permission.)
Figure 3.8 Conventional BRDF plot of perfect mirror (“instrument”) compared with a
perfect diffuse reflector and two mirrors with scatter.
Figure 3.9 Special parameter versus BRDF plot of perfect mirror (“instrument”) compared
with a perfect diffuse reflector and two mirrors with scatter.
3.7 Emission
So far, we have considered the radiometric properties of materials with respect to
incoming radiation. In fact, all materials above 0 K radiate, so the emission of
radiation by a material is an important property, as well. The Infrared Handbook
defines emissivity as “the ratio of the radiant exitance or radiance of a given body
to that of a blackbody.”10 Its symbol is ε. Emissivity may be considered a
“quality” factor, indicating the capability for thermal radiation by a material. It
has both spectral and directional properties, it is dimensionless, and its values are
between 0 and 1. As with the material properties already discussed, emissivity
refers to the characteristics of a pure substance, while emittance refers to the
properties of a specific sample.
Spectral emittance ε(λ) is defined as emittance at a given wavelength, and it
is not a derivative quantity. In the case that a radiator is neutral with respect to
wavelength, with a constant spectral emittance less than 1, it is called a graybody.
In that case, the spectral emittance is the ratio of the radiance of that source at
that wavelength to the radiance of a blackbody at that wavelength:
Lλ
ε(λ ) = , (3.23)
LλBB
where LλΒΒ is the value of the Planck function at that wavelength for a
blackbody. Further discussion of this function and the radiometric characteristics
of sources follows in Chapter 4.
As reflectance, transmittance, and absorptance are related, as indicated by
Eq. (3.10), so too are reflectance, transmittance, and emittance related. At
equilibrium, the power emitted by a body to its surroundings must equal the
power absorbed by the body from its surroundings. More succinctly, the body’s
absorptance must equal its emittance:
α = ε. (3.24)
ε =1− τ − ρ , (3.25)
and spectrally,
ε =1− ρ , (3.27)
and
ε(λ) = 1 − ρ(λ ) . (3.28)
These materials are often used for temperature control of spacecraft. Heating
is by absorption of sunlight for wavelengths shorter than 3 μm, and cooling
results from thermal radiation for wavelengths longer than 3 μm. If a material has
high reflectance at shorter wavelengths, it will not absorb much of the incident
radiation. If the reflectance is low at longer wavelengths, the absorption and
consequently the thermal emission will be high. The surface will be cold. If the
spectral regions are reversed, the surface will become hot. These surfaces are
known as selective surfaces, and a wide range of surface temperatures have been
achieved.
Designers of these materials often utilize the ratio α/ε to describe the value of
absorptance in one spectral region (usually solar) relative to the emittance value
in another region (usually infrared.) Figure 3.10 shows the spectral emittance
(1 – spectral reflectance) for several generic surfaces; some are selective.
The Infrared Handbook provides detailed examples of the radiometric
properties of both natural and artificial sources. An example of the spectral
reflectances of several metals is shown in Fig. 3.11.
Figure 3.11 Spectral reflectance of films of silver, gold, aluminum, copper, rhodium, and
12
titanium.
Optical properties
Transmission (function of wavelength, temperature,
direction)
Index of refraction (function of wavelength,
temperature, direction)
Dispersion, partial dispersion
Surface reflectance
Scatter (surface & bulk)
Absorption (bulk)
Homogeneity
Birefringence, stress coefficient
Fluorescence
Anisotropy
Electro-optic and/or acousto-optic coefficients
Mechanical properties
Young’s modulus
Yield point
Hardness
Optical workability
Coating compatibility
Density, specific gravity
Thermal properties
Thermal conductivity
Specific heat, heat capacity
Coefficient of linear thermal expansion
Softening point, melting point
Environmental properties
Solubility in H2O, other solvents
Surface deterioration, devitrification
Radiation susceptibility (UV, hard particle)
Other factors
Availability
Safety factors, toxicity
Cost
Compiled by James M. Palmer 02/21/89
References
4.1 Introduction
From a discussion in the previous chapter on the interaction of radiation with
materials, we now turn to the subject of how radiation is generated, and the roles
that emission, reflection, and other processes play.
Sources of optical radiation can be classified in a variety of ways. Active
sources emit optical radiation due to their temperature (thermal sources) or as a
result of atomic transitions (luminescent sources). Passive sources reflect optical
radiation from active sources or from other passive sources. Passive sources can
also be classified as thermal or luminescent, depending upon the process that
generated the radiation initially. Examples of thermal sources include blackbody
radiation simulators, tungsten-filament lamps, gases, the sun, the moon, and you
and I. Examples of luminescent sources include lasers, fluorescent lamps,
mercury arcs, sodium lamps, electroluminescent panels, LEDs, and gases. Some
sources combine both thermal and luminescent mechanisms, and some may be
both active and passive, reflecting in one spectral region and emitting in another.
Other means of classification have also been used. Some authors distinguish
between artificial (man-made) and natural sources. Lamps are artificial sources,
whereas the earth, the sun, and stars are natural sources. Still another practice is
to divide sources according to their output spectral characteristics. Continuous
sources have a spectral radiance that is slowly varying with wavelength, typical
of thermal radiation, while line sources emit in narrow, well-defined spectral
regions. Yet another attempt to distinguish sources is by their degree of spatial
and/or temporal coherence.
Thermal radiation has been extensively studied since the late nineteenth
century. Stefan was the first to experimentally examine the relationship between
radiation and temperature in 1879. He analyzed data from Tyndall and found that
the total radiation is proportional to the fourth power of temperature T4.
Boltzmann derived this T 4 relationship from the Carnot cycle in 1884. In 1891
Wien derived his displacement law, which relates the peak radiation to
83
temperature, and the T 5 relationship for the magnitude of the peak. In 1896, Wien
derived an equation for the spectral distribution of thermal radiation based on
thermodynamic arguments. In 1900 Rayleigh derived another equation for
spectral distribution based on equipartition, and Jeans in 1905 independently
repeated this derivation. Planck in 1901 published an empirical equation
involving the notion that energy exists as discreet “packets” that fit experimental
data better than either the Wien or the Rayleigh-Jeans equation. This proved
monumental, as it was later accepted as the birth of quantum mechanics.
Verifications of the Planck equation and the physical constants continued up to
1982 with precise measurements of the spectral distribution of blackbody
radiation, determination of the Stefan-Boltzmann constant, and confirmation of
the thermodynamic temperature scale.
8πν 2
N ν dν = dν . (4.1)
c3
hν
q= hν / kT
. (4.2)
e −1
The energy density (energy per unit volume per unit frequency interval) is
the product of the number of available modes and the average energy of each
mode. It is expressed as:
8πhν3 1
uν = . (4.3)
c 3 (e hν / kT − 1)
The energy exits a cavity at velocity c into 4π sr. The behavior is described by
Planck’s law, expressing radiance per unit frequency interval (spectral radiance):
2hν 3 1
Lν = 2 hv / kT
. (4.4)
c e −1
Lλ d λ = Lν d ν
and the relationship between wavelength and frequency λ = c/ν, the spectral
radiance of blackbody radiation can be expressed in wavelength terms as
2hc 2 1
Lλ = 5 hc / λkT
. (4.5)
λ e −1
c1 1
Lλ = 2 5 c 2 / nλT
. (4.6a)
πn λ e −1
c1 1
Mλ = 2 5 c 2 / nλT
. (4.6b)
nλ e −1
The two curves in Fig. 4.1 show the form of Planck’s equation as a function
of wavelength, with temperature as a parameter. They are strongly peaked, with
the form governed by the λ–5 term for wavelengths longer than the peak and by
the exponential term for shorter wavelengths. Note that only a limited range of
temperatures can be shown on a single linear plot such as these, as the ordinates
are highly nonlinear, varying over many orders of magnitude. The plots also
show a dashed line, the locus of the wavelength of peak spectral radiance, having
a characteristic hyperbolic shape. As will be seen, these dashed lines represent
the Wien displacement law.
If the spectral radiance curves are plotted on log-log axes as in Fig. 4.2,
several interesting things are seen. First, this form allows for a wide range of
temperatures and wavelengths on a single plot. Second, the locus of the
wavelength of peak radiance is a straight line on a logarithmic plot, indicating
hyperbolic behavior. Finally, note that all of the curves have an identical shape
when the logs are plotted. Since the shape of the curve is independent of
temperature, one could construct a nomogram by tracing a single curve and the
straight line locus of maxima onto a transparent sheet and use it as an overlay.
Slide the overlay along the straight line to display the blackbody radiation curve
for any temperature.
A rough but useful approximation for the peak wavelength (in μm, setting n = 1)
is
Table 4.1 lists the peak wavelengths for several common sources.
(a)
(b)
Figure 4.1 Spectral radiance of blackbody radiation for (a) high temperatures and (b)
lower temperatures.
Figure 4.2 Log-log plot of blackbody spectral radiance as a function of wavelength and
temperature.
To find the value of Lλ at the peak wavelength, solve the blackbody equation
using the peak wavelength (in the medium):
n 2 σ′ 5
Lλ max = T , (4.9)
π
where σ' = 1.286 × 10–11 W/m2K5μm. The radiance at the peak wavelength varies
as the fifth power of the temperature.
∞
n2σ 4
L = Lλ d λ = T , (4.10)
0
π
M = σT 4 . (4.11)
c
Q = hν = h , (4.12)
λo
*
More accurate values for the constants that appear in this section are given in Appendix B; higher
accuracy is sometimes required.
c1q 1
Lqλ = 4 c2 / λT
, (4.13)
πλ e −1
where c1q = 2πc = 1.883 × 109 m/s. For wavelength in μm, c1q becomes 1.883 ×
1027 m2/μm·s. Thus, Lqλ is expressed in photons per second per area per unit
wavelength.
Plots of spectral photon radiance are shown in Figs. 4.3 and 4.4. The curves
appear similar to the previous radiance curves, but the range of ordinate values is
not so extreme. The equation is subjectively the same; the exponential term is
identical but the wavelength in the denominator is only raised to the fourth
power.
(a)
(b)
Figure 4.3 Photon spectral radiance versus wavelength for (a) high temperatures and (b)
lower temperatures.
Figure 4.4 Log-log plot of spectral photon radiance as a function of wavelength and
temperature.
The Wien displacement law for photons is derived as described earlier; the
result is
σ′q
Lqλ max = T4 , (4.16)
π
∞
σq
Lq = Lqλ d λ = T3, (4.17)
0
π
c1 1
Lλ = 5 c 2 / λT
.
πλ e −1
c 2 / λT c2 (c2 / λT ) 2 (c2 / λT )3
e =1+ + + + ... .
λT 2! 3!
If c2/λT << 1 (corresponding to a large value of λT), drop all of the higher-order
terms. Then,
c 2 / λT c2
e −1 ≈ .
λT
Rearranging terms and substituting into the Planck equation, the result becomes:
2ckT
Lλ = . (4.18)
λ4
−c
c1 λT2
Lλ = e (4.19)
πλ 5
The Wien approximation is valid with less than 1% error if λT < 3000 μm·K
(short wavelengths and/or low temperatures). It is quite useful for a great deal of
radiometric work as it is valid for blackbody radiation at all wavelengths shorter
than the peak. Figures 4.5 and 4.6 depict the curves and the ranges of validity for
the two approximations.
Lλ c1 1
= . (4.20)
T 5 π(λT )5 ec2 / λT − 1
(μm·K)
(μm-K)
Figure 4.6 Errors associated with Wien and Rayleigh-Jeans approximations.
Lλ (λT ) L (λT )
f (λ T ) = = λ (4.21)
Lλ (λ maxT ) σ' 5
T
π
and
λ λ
Lλ (λT )d λ L (λT ) d λ
λ
F (λT ) = 0
= 0
. (4.22)
∞
σ 4
L (λT ) d λ
0
λ
π
T
5E-12
4E-12
RADIANCE/T^5
3E-12
2E-12
1E-12
0
0 5000 10000 15000
WAVELENGTH x TEMPERATURE (um-K)
Since these definitions are ratios, exitance M can readily be substituted for
radiance L. These two functions are graphed in Fig. 4.8.
To use these curves to determine radiance in a narrow wavelength interval
(Δλ < 0.05λc), first select T for the blackbody radiation and the desired center
wavelength λc. Determine the radiance at the peak using the equation for Lλ(max).
Finally, use the function f(λT) from the graph and the wavelength interval Δλ to
arrive at the result
λ c +Δλ / 2
λ c −Δλ / 2
Lλ (λT )d λ = Lλ (λ maxT ) f (λ cT )Δλ . (4.23)
If the wavelength interval is large, typically greater than 0.05× the center
wavelength, use the other function F(λT) to determine the radiance in a finite
wavelength interval. Again select T and the two desired wavelengths, λ1 and
λ2. From the graph, read F(λ1T) and F(λ2T) and compute the total radiance using
the Stefan-Boltzmann law. The result is
λ2
σ
L (λT ) d λ = π T [ F (λ T ) − F (λ T ) ]
4
λ 2 1 λ2 > λ1. (4.24)
λ1
Figure 4.8 also shows the corresponding curves for photons, fq(λT) and
Fq(λT). The defining equations are
Figure 4.8 Curves of f(λ) and F(λ) for watts and photons.
Lqλ ( λT )
f q ( λT ) = (4.25)
Lqλ ( λ maxT )
and
λ λ
Lqλ (λT )d λ L qλ (λ T ) d λ
Fq (λT ) = 0
= 0
. (4.26)
∞
σq
L
3
qλ (λT ) d λ T
0
π
The application of the fq and Fq photon curves is identical to the curves for
energy.
In some applications, it is desirable to maximize the radiation contrast
between a target and background of similar temperature. What wavelength might
one choose for this task? The problem occurs regularly in the infrared where both
the target and the background radiate near 300 K. Take the second derivative
d2Lλ/(dTdλ), and set it to zero. The result is:
This equation implies that the best “visibility,” or contrast with the
background, occurs at a wavelength somewhat shorter than the peak wavelength,
at λcontrast = 0.832λmax. Since the result is on the short wavelength side of the
peak, the Wien approximation is valid and it simplifies the calculus considerably.
Thus, if your target and background temperature were 305 K, the peak
wavelength is 9.50 μm and the wavelength of maximum radiation contrast λcontrast
is 7.9 μm. The wavelength for maximum photon contrast is 2898 μm·K, the same
as the peak wavelength for energy.
We often need to know how the spectral radiance Lλ changes with
temperature; this can be determined by differentiating the Planck function with
respect to temperature. The result, shown in differential form, is
ΔLλ xe x ΔT
= x , (4.28)
Lλ e −1 T
hc c
where x = = 2 .
λkT λT
xe x
We define Z ≡ x . For a small change in temperature ΔT such that the
e −1
change in x is also small, the change in Lλ with temperature at any wavelength is
ΔLλ ΔT
=Z . (4.29)
Lλ T
Lλ (θ, φ)
ε(λ; θ, φ) = . (4.30)
LλBB
This equation differs slightly from Eq. (3.23), as it includes the directional
component.
Total emissivity is the integral of spectral directional emissivity over all
angles and wavelengths, also weighted by the Planck function, which introduces
a temperature dependence if the spectral emissivity is not uniform (that is, gray):
ε=
(θ, φ; λ) L λBB sin θ cos θd θd φd λ
. (4.31)
(σ / π)T 4
ε ≠ (θ, φ; λ )d λ . (4.32)
λ
α = ε. (4.33)
by the body to the enclosure must equal the power absorbed by the body from the
enclosure:
αΦ i = εΦ BB .
As the incident and emitted power are the same (no other sources or sinks), the
conclusion is α = ε.
This seems too simple; is it really true? No! Look at the following equation
for the equilibrium temperature of a flat plate in space, facing the sun and
insulated on the back side. It is a simple statement of conservation of power,
equating radiant incidence (absorbed) to radiant exitance (emitted):
Here Eo is the solar constant (1368 W/m2). The units on both sides of the
equation are W/m2. Solving for T, we find that it is a function of the ratio (α/ε).
But Kirchhoff’s law says that α = ε, and these terms therefore cancel. The
equilibrium temperature is therefore a function only of Eo. The implication is that
a white (allegedly reflective) car and a black (supposedly absorptive) car have the
same equilibrium temperature after sitting out in the sun all day. People from
Arizona and Florida know better than that, and often buy white cars to keep
cooler in the summer! So what’s wrong?
The answer is that Kirchhoff's law does not apply in all situations. Table 4.2
indicates the applicability in terms of spectral and directional conditions.
Specifically, α(λ;θ,φ) = ε(λ;θ,φ); the absorptance equals the emittance at a single
wavelength in a single direction. After integrating over wavelength and
geometry, test for the stated restrictions before applying Kirchhoff’s law. The
1
Figure 4.10 Illustration of Kirchhoff’s law. [Reprinted with permission of author from
Optical Radiation Measurement series, Vol. 1, F. Grum and R. J. Becherer, Radiometry, p.
98 (1979).]
simple equation above attempts to relate the absorbed energy from the sun, which
is concentrated in the spectral range between 0.3 and 3 μm, to the thermal
radiation located in the wavelength range between 3 and 30 μm. White and black
paints have completely different reflectances in the solar region, but their
emittances at 300 K are very similar.
The thermal radiative properties of any material can be given in terms of its
temperature T and its spectral directional emittance ε(λ;θ,φ). The spectral
dependency is of primary importance, and the directional properties and
temperature coefficients are usually of lesser concern.
The blackbody equation is highly nonlinear with both wavelength and
temperature, and is thus not particularly tractable. Fortunately, there are many
computational and visual aids to help. The Infrared Handbook gives several
calculator programs for now-obsolete HP and TI calculators, and some early
calculators were available with plug-in cards to do blackbody calculations. The
latest incarnation, The Infrared and Electro-Optical Systems Handbook,
substitutes an extensive set of programs in BASIC to do the computations. There
have been several slide rules (remember them?) that do calculations on
blackbody radiation. The GE Radiation Calculator is a plastic rule that
occasionally surfaces at a reasonable price. Cardboard knock-offs are often given
away by vendors. For a little more money, Electro-Optical Industries at one time
sold a high-quality metal rule. One side of my venerable GE calculator is shown
in Fig. 4.11.
Back when computers were scarce, tables generated by mainframes were
commonly used to do precise blackbody calculations. Several tables may still be
found in musty libraries.2,3,4
Today, simple computer programs can easily be written. Spreadsheets are
sufficiently powerful to do these calculations with relative ease and provide
superior graphics as well. Tools like Mathcad, MATLAB, and Mathematica also
work well.
1
Table 4.2 Summary of absorptance-emittance relations. [Adapted with permission of
author from Optical Radiation Measurement series, Vol. 1, F. Grum and R. J. Becherer,
Radiometry, p. 115 (1979).]
Required conditions:
Required conditions:
(1) Incident energy independent of angle and spectral distribution proportional to
blackbody at T, or
(2) Incident energy independent of angle and α(λ;θ,φ,Τ )
= ε(λ;θ,φ,Τ ) independent of wavelength, or
(3) Incident energy at each angle has spectral distribution proportional to
blackbody at T and α(λ;θ,φ,Τ )
= ε(λ;θ,φ,Τ ) independent of angle, or
(4) α(λ;θ,φ,Τ ) = ε(λ;θ,φ,T ) independent of angle and wavelength
Spectral emittance was defined as the ratio between the spectral radiance of
an object and the spectral radiance of a blackbody at the same wavelength and
temperature. In general, it is also a function of angles θ and φ. Using this
definition, any isothermal object has a spectral radiance given by
ε(λ; θ, φ)c1 1
Lλ (θ, φ) = 5 c2 / λT , (4.35)
πλ e −1
and if ε(λ;θ,φ) and T are known, so is the spectral radiance. As ε(λ) increases
towards unity, blackbody radiation is approached. Blackbody radiation
simulators attempt to do just this, and many are commercially available. Since
the spectral radiant exitance can be described by this simple equation, blackbody
radiation simulators are most often used as standard sources of optical radiation
in the calibration laboratory.
4.3.1 Metals
An important factor in the behavior of a material is its electrical conductivity.
Metals have a high conductivity, meaning large quantities of free electrons are
available to interact directly with the radiation field.
Since metals are generally good reflectors (a direct consequence of their high
conductivity), they are poor emitters (conservation of energy). The emittance is a
slowly varying function of wavelength, as shown by a simple empirical equation
given by Hagen and Rubens:5
resistivity
ε ≈ constant (λ > 2 μm) (4.36)
λ
This equation, which is generally valid at wavelengths longer than 2 μm, shows
that increasing resistivity (decreasing conductivity) increases emittance, and that
emittance decreases at longer wavelengths. Figure 4.12 gives examples.
The directional properties can be derived from Maxwell’s equations. The
radiation is highly polarized at angles off normal, as shown in Fig. 4.13. Note,
however, that the total radiation, the sum of both polarizations, is quite constant
with angle (i.e., Lambertian) to within a few percent out to nearly 60 deg from
specular.
4.3.2 Dielectrics
Dielectrics and gases have much lower conductivity than metals. Their electrons
are more tightly bound to their parent nuclei and require specific atomic
interactions with the radiation field. This implies that dielectrics tend to radiate in
specific, fairly well-defined spectral regions, and not elsewhere.
The emitting properties of dielectrics are closely associated with the complex
index of refraction n + iκ where κ, the extinction coefficient, is in turn related to
the absorption coefficient α′, as discussed in Chapter 3. As noted in that chapter,
the product of the absorption coefficient and the thickness of a material is the
optical thickness τo. A material is considered optically thin when τo < 0.1
(transmission high) and optically thick when τo > 2 (transmission low). Optically
thin materials approach transparency and have low emittance; for optically thick
materials, the normal emittance is (1 – reflectance) and depends on the index of
refraction as determined by the Fresnel equation at normal incidence as
illustrated in Fig. 4.14. Emittance at other angles also comes from the Fresnel
equations and is polarized.
4.3.3 Gases
Gases are optically thin over wide wavelength ranges and may be transparent
over long paths. Therefore, their emittance is essentially zero at these
wavelengths. However, there are specific spectral regions where absorption, and
therefore emission, occur. Each species has its own absorption characteristics,
correlated with its atomic and molecular structure and energy levels. These
characteristics take the form of a series of spectral lines at regular locations in the
spectrum. They are occasionally seen as discrete lines, but more often as a series
of overlapping lines called bands.
The important species identified with our own atmosphere include “fixed”
gases O2, N2, and CO2. Those species classified as variable include water, ozone,
and methane (H2O, O3, and CH4.) It must be stressed that the energy absorbed by
a gas is dependent upon the concentration and absorbing characteristics of the
gas, and the energy emitted by the gas depends upon the temperature and spectral
emittance of the gas. The processes of absorption and emission are not
independent, but occur simultaneously.
0.9
0.8
0.7
0.6
1 2 3 4
INDEX of REFRACTION
Fig. 4.15. He measured the emittance of a flat ribbon of tungsten, but the results
are also applicable to a round wire. Most lamps are made by drawing the
tungsten into a round wire and then tightly coiling it, enhancing the emittance by
making a partial cavity. If this tight coil is then further loosely coiled, emissivity
is enhanced further.
Tungsten lamps are designed to operate at a nominal voltage and current. The
design compromise is between light and lifetime. If more light is needed, voltage
(or current) may be increased, but not for very long. Figure 4.16 shows light
(lumens), efficiency (lumens/watts), and lifetime as a function of operating
voltage.
Tungsten-filament lamps decay due to evaporation of the filament, leaving
brown deposits on the inside of the lamp envelope. An uneven rate of
evaporation creates “hot spots” which cause the lamp’s overall rate of decay to
increase. Additionally, filaments may crystallize and become brittle particularly
when the lamp is operated on dc, mechanically weakening the filament and
making it susceptible to breakage from mechanical or thermal shock.
Tungsten-halogen lamps are better suited than tungsten for most applications,
as they either eliminate or delay the onset of both decay mechanisms. The
addition of a halogen such as bromine or iodine creates a regenerative cycle, in
which the evaporated tungsten combines with the halogen rather than plating on
the envelope. A hot (minimum 250° C) envelope, usually of fused silica, is
required. The resulting halide compound decomposes at a rate proportional to
temperature. This decomposition occurs preferentially at the “hot spot,” causing
tungsten to plate back onto the filament.
Figure 4.15 Spectral emittance of tungsten. (Reprinted from Ref. 7 with permission from
Elsevier.)
are being phased out in the European Union, Australia, the United States, and
other nations.
200 W and operates at 1500 to 2000 K. The emittance is about 0.75 out to a
useful wavelength of 30 μm. It is rather fragile and has a lifetime of 200 to 1000
hours. Its resistance at room temperature is extremely high, requiring an auxiliary
platinum heater coil to get it started. Nernst glowers are most often found in
infrared spectrometers where their shape is ideally suited to image onto a narrow
slit.
A popular low-cost IR source consists of an aluminum oxide (Al2O3) ceramic
tube heated with a coaxial nichrome wire. This source is often used in
inexpensive infrared spectrophotometers and operates at about 1200 K with an
emittance of about 0.8. The power source is a simple line-operated transformer.
The Welsbach mantle is a woven fabric mesh impregnated with refractory
oxides such as thorium oxide. It is heated with propane or white gas to a
temperature up to 2400 K. The emittance is high for λ > 10 μm. This source is
commonly known as the “Coleman” lantern.
The carbon arc is a valuable source for announcing grand openings, detecting
hostile aircraft, simulating solar radiation, etc. The radiation comes from the
plasma-heated carbon at its sublimation temperature of 3800 K. By placing
refractory oxides in the carbon rod, higher temperatures commensurate with the
evaporating points of these oxides can be achieved, in the range 5000 to 8000 K.
The high brightness comes from the small (about 10-mm diameter) size of the
carbons. The disadvantages are lack of stability, mechanical issues involving
continuous feeding of carbons, the power required (usually a noisy, smelly,
unregulated motor generator) and the need for ventilation (hydrogen cyanide is
generated).
like the hot soot from burning hydrocarbons with insufficient oxygen. Examples
include the oxyacetylene torch before the oxygen is turned on, and flames from
solid rocket motors. These particles are effectively “black,” emitting as small
blackbody radiators. The other component is radiation from the gases, including
impurities. For example, the yellow color from a Bunsen burner is due to atomic
emission from sodium in the supply gas.
Exhaust gases emit radiation in spectral regions where they are optically
thick. The important gases (from a radiometric standpoint) are the hot
combustion products CO, C2O, and H2O. Particulates, “black” if their size is
> 30 μm, are often present, particularly from diesel exhaust and muzzle flash.
If a gas both emits and absorbs at the same wavelength, how can we see it?
Our initial reaction is that the absorption lines from these gases block the
emission from being seen. In fact this is not the case. A phenomenon known as
“line reversal” takes place wherein hot gases have slightly different spectral
profiles than their unheated counterparts, broader because of increased
temperature and pressure. Absorption will occur at the center wavelength of the
profile, but the hot gas may be observed in two wavelength bands bordering the
center.
You and I make good infrared sources. Our temperature is about 300 K and
our emittance is nearly unity for all wavelengths longer than about 5 μm. We are
all black in the thermal infrared, regardless of skin pigmentation in the visible.
Appliances and conveyances also make interesting sources. Cars, trucks,
trains, tanks, and aircraft all have different temperatures than their surroundings
when at work. Even such little things as insulators on a power transmission line
get hot when leakage occurs, rendering them observable targets in the infrared.
Table 4.4 lists the total (integrated over wavelength) directional (normal
incidence) emittance for several materials at the indicated temperatures (in
degrees kelvin).
At low pressures (below 100 mbar), the Doppler effect predominates. The
shape of the curve is Gaussian, and a typical equation takes the form:
( v − vo )2
0.47 S − ln 2 α2
k (ν ) = e , (4.37)
α
where k(ν) is an extinction coefficient at frequency ν, S is a line strength, α is an
absorption coefficient related to the halfwidth of the line, and νo is the line
Sα
k (ν ) = . (4.38)
π (ν − ν o ) 2 + α 2
Doppler
Lorentz
8
Figure 4.18 Emission lines for Hg (< 600 nm) and Ar (> 600 nm). (Copyright held by
Ocean Optics, Inc. Reproduced with permission.)
Figure 4.21 Components of the compact fluorescent lamp. (Reprinted with permission
th
from the IESNA Lighting Handbook, 9 Edition, by the Illuminating Engineering Society of
10
North America.)
4.4.3.5 Lasers
The LASER (light amplification by stimulated emission of radiation), was first
demonstrated in 1960, and represented a radical departure from conventional
sources known at the time. The combination of stimulated (rather than
spontaneous) emission to provide gain and a frequency-selective feedback
mechanism gives the laser its unique properties. The stimulated emission results
from a population inversion, where there are more electrons at a higher energy
level than at a lower energy level. This population inversion can be generated
using one of several different energy sources, including electrical, optical,
magnetic, chemical, or nuclear. The frequency-selective feedback mechanism
most often takes the form of a resonant cavity with mirrors. For frequencies
having overall gain greater than unity, oscillation is possible.
Lasers differ from incoherent sources in many ways. Since the amplification
process depends upon strict phase relationships, the radiation produced is
coherent, i.e., the waves are in phase with each other. Since we are dealing with
atomic transitions and a resonant cavity, the output is very nearly
monochromatic, with a very narrow bandwidth. Because of the requirements
placed on the resonant cavity, the beam diameter and the beam divergence angle
are both small.
Lasers come in all forms, shapes, and sizes. Some generate unmodulated
light [continuous wave (CW)] with power levels from milliwatt to megawatt and
above, while others generate pulses. Pulse widths can vary from several
femtoseconds (10–15 s) to several milliseconds with repetition rates varying from
gigahertz to millihertz. Size can be as small as a TO-18 transistor to as large as a
full-size laboratory.
A wide variety of materials have been found useful as gain media. Gas lasers
include He-Ne and CO2. Examples of ion lasers are argon and krypton. Solid-
state lasers are exemplified by ruby and Nd:YAG. A semiconductor laser is
similar to an LED but incorporates the resonant cavity needed for laser operation
within the semiconductor structure. Tunable lasers are available using organic
dyes and special solid-state crystals such as alexandrite. Erbium can be doped
into a fiber to provide in situ gain.
4.4.4.1 Sunlight
As Table 4.3 shows, natural sources appear in both thermal and luminescent
categories.
The most prominent natural source of radiation, indeed the most important, is
our sun. The best estimate for the solar constant is 1368 W/m2 with an
uncertainty and drift of about 0.2%. Since the orbit of the earth around the sun is
elliptical, there is an additional diurnal variation of ±3.5%, with the maximum
experienced in January.
Things are somewhat different at the surface of the earth. Our atmosphere
selectively attenuates by two primary means: scattering by molecules (Rayleigh
scatter) and aerosols (Mie scatter), and absorption by molecules (H2O, CO2, and
O3). The term “air mass” is frequently used to indicate how much atmosphere the
solar radiation is traversing; it is approximately equal to the secant of the solar
zenith angle.† Figure 4.23 shows extraterrestrial solar spectral irradiance in the
visible and near infrared, along with transmitted solar spectral irradiance for
several air masses.
†
At or near sea level. As the altitude increases significantly, and air pressure goes down, the air
mass decreases significantly.
12
Figure 4.23 Spectral characteristics of direct sunlight.
Figure 4.24 Irradiance from skylight.13 [Reprinted from Daylight and its Spectrum, S. T.
Henderson, p. 113 (1970).].
Figure 4.25 Earth radiance in the thermal infrared region. Dashed curve: blackbody at 272
K; solid curve: radiance of earth from space; dotted curve: radiance seen from earth’s
14
surface, looking up.
Ultraviolet eye damage is possible with UV sources; shield the lamp and wear
protective goggles.
There is an explosion hazard from high-pressure lamps. Use an approved
housing, clean quartz envelopes thoroughly, handle lamps with gloves, and
wear eye protection.
A fire hazard exists with powerful lamps; keep flammable materials away from
them.
Ozone is generated by ultraviolet lamps; use ozone-free lamps (envelopes that do
not transmit UV) or provide proper ventilation.
There are electrical hazards from lamp power supplies; take normal precautions,
particularly with high-voltage starters for arc sources and capacitor banks for
pulsed lasers.
Careful handling and disposal of CFLs is a must! See http://www.energystar.gov
for more information.
ε(λ; θ, φ)c1 1
Lλ (θ, φ) = 5 c2 / λT .
πλ e − 1
References
5.1 Introduction
Optical radiation detectors are transducers that transform optical radiant energy
into a different form of energy that is more readily measured. Electrical energy is
typically used for this purpose, as electrical measurement technologies are well
established. Both thermal and photon detectors convert incident optical energy
into electrical signals; in the thermal detector, the initial output takes the form of
heat before conversion. Either detector type may be a “point” or an “area”
detector. The former are single-element detectors, designed to respond to incident
energy. The latter are one- or two-dimensional arrays used particularly for
imaging, and include mechanisms to read out the signal on the array. Table 5.1
gives examples of photon and thermal detectors, while Table 5.2 lists differences
between them.
Thermal Photon
Low detectivity High detectivity
Slow response time Fast response time
Do not require cooling Typically require cooling for
IR operation
127
Figure 5.1 Characteristic response curves of (a) photon and (b) thermal detectors.1
[Reprinted from Optical Radiation Measurement series, Vol. 1, F. Grum and R. J.
Becherer, Radiometry, p. 177 (1979).].
The parameters in Table 5.2 will be discussed in more detail later in this
chapter. Figure 5.1 illustrates the generic differences in spectral response
between photon and thermal detectors. Note that the thermal detector response is
wavelength independent, and that the photon detector response is a function of
several parameters including wavelength and quantum efficiency. A specific
example of the latter is the commonly used silicon detector, whose wavelength of
peak response is 950 nm and whose cutoff wavelength is near 1100 nm.
5.2 Definitions
Several terms commonly used to describe detector parameters are not often used
outside the field of optics. Recall that the term wavelength refers to the optical
regime below 100 μm. The term frequency is used to describe the direct current
(dc) to 1012 Hz audio/radio regime. The following defintions will aid our study of
detectors.
Spectral responsivity ℜ(λ) is the ratio of the output of a detector or
radiometer to that of a monochromatic source of optical radiant power. The
detector ouput is typically a current (amperes, A) or a voltage (volts, V), while
the incoming optical quantity is power, measured in watts. It is also a function of
wavelength, in that it is measured at specific wavelengths, and is correctly
reported in A/W or V/W at a specific wavelength. It can also vary as a function
of detector temperature and input power level.
Responsivity ℜ is the ratio of the output of a detector or radiometer to the
incoming optical radiant power, integrated over the spectral range which is
common to the source and detector. The input and output signals are those
described above. Since this parameter is the result of the quotient of two integrals
as shown below, it is source dependent for nonmonochromatic or nonflat sources.
ℜ=
R (λ )Φ (λ ) d λ . (5.1)
Φ (λ ) d λ
On occasion, one will also find responsivities related to other radiometric
quantities, such as irradiance responsivity ℜE in amps (or volts) per W/m2, or
radiance responsivity ℜL in amps (or volts) per W/m2sr. These responsivities are
more likely to characterize radiometric instruments than detectors alone.
Photon spectral responsivity ℜq(λ) is the ratio of the output of a detector or
radiometer to the monochromatic optical radiant photon flux incident upon the
detector or radiometer. The output is typically a current (A) or a voltage (V),
while the incoming optical quantity is photon flux (photons/s). Therefore
ℜq(λ) carries the units (A·s) or (V·s). This parameter is wavelength dependent for
most detectors and also can vary with temperature and input power level. Like
spectral responsivity, photon spectral responsivity is a function of wavelength.
Photon responsivity ℜq is the ratio of the output of a detector or radiometer
to the incoming photon flux, integrated over the spectral range which is common
to the source and detector. The input and output signals are those described
above. Since this parameter is the result of the quotient of two integrals, it is also
source dependent for nonmonochromatic or nonflat radiation sources.
The electrical signal that is output by a detector is governed by the electrical
characteristics of the detector (resistance, capacitance, etc.) as well as its
associated circuitry. A Bode plot is a convenient way to depict a system’s
electrical frequency response, with gain or phase plotted on the ordinate and
frequency on the abscissa. For our purposes, voltage or current gain are often
depicted. Both axes are logarithmic, as seen from the generalized Bode plot in
Fig. 5.2. With the proper choice of axes, slopes associated with single time
constants (6 dB/octave, 20 dB/decade) plot as 45-deg lines, and the cut-on and/or
cutoff frequencies are at the intercepts of the straight-line asymptotes.
Time constant τ is the time required for a signal to achieve 63% of its final
output, given a step input. For a simple single-section resistance-capacitance
(RC) circuit, τ = RC. The rise time is the time the signal takes to rise from 10%
of its final value to 90%. For a single time-constant circuit, its value is
2.2τ. Figure 5.3 illustrates this concept.
Cutoff frequency fc is the frequency at which the voltage or current response
of a detector or circuit falls to 1 2 , or 0.707× its dc or midband maximum value.
It is related to the time constant by fc = 1/(2πτ). In terms of power, it is the
frequency where the power drops to half of the dc or midband maximum value.
This is also called the 3-dB frequency.
The root-mean-square (rms) value of a quantity (voltage in this example) is
defined as
T
1 2
T 0
vrms = v (t )dt , (5.2)
where T is a single or integer multiple period for periodic waveforms. The rms
values discussed in this chapter may be values of voltages or currents and are
often referenced in descriptions of noise.
Signal is the component of the output voltage or current from a detector that
arises from a specific radiometric input. Signal is what you want to work with. It
is an rms current or voltage and is denoted vs or is. The signal is the integral of
the product of the spectral responsivity and the radiometric input such that:
∞
SIGNAL = R(λ)Φ (λ)d λ . (5.3)
0
RISE TIME
1
0.8
0.6
OUTPUT
0.4
0.2
0
0 0.5 1 1.5 2 2.5 3
TIME (time constants)
Noise is the component of the output voltage or current from a detector that
arises from random fluctuations in the detector circuit, in the incoming radiation,
or from a number of other sources discussed in this chapter. It is characterized by
an rms current or voltage and denoted vn or in.
Signal-to-noise ratio (SNR) is the ratio of the rms signal current is to the rms
noise current in. Voltage may be substituted for current when calculating SNR,
which is dimensionless.
Impedance Z is the slope of the current versus the voltage curve of a device
at the designated operating point:
dV
Z= (οhms) . (5.4)
dI
SNR 2 out
DQE = . (5.5)
SNR 2 in
in i
NEP = = Φ n ; or
ℜ is
(5.6)
vn vn
NEP = = Φ .
ℜ vs
Note that NEP is a useful figure of merit only when the limiting noise is
inherent in the detector, and not due to the input signal.
Spectral NEP(λ) is the NEP for a monochromatic single-wavelength input
signal. Again, this is not a derivative quantity, per-unit wavelength, but merely a
value at a specific wavelength.
Noise-equivalent photon flux NEΦq is the incoming signal in photons/s that
produces a signal-to-noise ratio of unity. It is therefore the ratio of the rms noise
current (or voltage) to the photon responsivity.
Detectivity D is the reciprocal of NEP, originally defined as a term because
“bigger is better.” The unit is W–1. It is conveniently thought of as the SNR for a
1-W input. Multiply D by the input power to get the SNR. Like responsivity ℜ
and NEP above, this term is a result of integration over wavelength:
1 SNR
D= = . (5.7)
NEP Φinput
Ad B Ad B vs
D* = D Ad B = = cm·Hz1/2/W. (5.8)
NEP Φ vn
In the expression above, signal and noise currents may replace signal and
noise voltages. To obtain the SNR from D*, multiply D* by the input power and
divide by (AdB)1/2.
This term is called blackbody D* if the incident power comes from a
blackbody radiation simulator. The notation used in this case is D*(T,fc,B), where
T is the blackbody radiation simulator temperature in degrees kelvin (500 K is
common), fc is the chopping frequency in Hz, and B is the effective noise
bandwidth, by convention, 1 Hz.
Spectral D*(λ) is D* at a specific wavelength. It is defined as
Ad B
D * (λ ) = cm·Hz1/2/W. (5.9)
NEP (λ)
Ω
D** = D* = D* sin Θ1 2 , (5.10)
π
hc AdB is
D *q = D * = cm·Hz1/2/(phot/s), (5.11)
λ n in
where n-bar in the denominator is the photon flux in photons per second.
BLIP is an acronym for background-limited infrared photodetector, the
condition in which the limiting noise in a detector output arises from background
photons. A BLIP detector’s internal noise has been reduced to the point where it
is not significant. This ideal condition allows SNR to be easily calculated.
RA product is the product of the detector resistance and area and is a
constant for many materials. D* is proportional to (RAd)1/2 for many photovoltaic
detectors; thus the RA product can be used as a figure of merit for material
comparisons.
5.4 #N$O%&I*S@E~^
Assuming that (2) and (3) can be minimized using good engineering
practices (shielding, grounding, proper component and circuit layout), we
concentrate on (1). Figure 5.4 is a plot of noise spectral density versus frequency
that includes several of the above noise mechanisms.
Noise comes in many forms and colors. The most prevalent model of noise is
white noise, having a power spectrum independent of frequency. Of course real
noise is never white, as that would imply infinite total power. Noise density
actually goes to zero above 1012 Hz because of the finite mass of electrons. In
Fig. 5.4, the region at frequencies greater than about 50 Hz is white
(superimposed on spikes from fixed frequency sources). Pink noise has a power
spectrum that increases with decreasing frequency at a nominal rate of 3 dB per
octave, proportional to the inverse of the frequency (1/f ). It occurs at low
frequencies, below about 5 Hz. Noise increasing at a rate of 6 dB per octave
(proportional to 1/f 2) at low frequencies is referred to as red (or brown) noise.
There are special names for relatively rare noises that increase with increasing
frequency. Blue noise has a power spectrum that increases 3 dB per octave with
increasing frequency (proportional to f ), and purple noise has a power spectrum
that increases 6 dB per octave with increasing frequency (proportional to f 2).
Figure 5.4 Noise spectral density as a function of frequency. (Reproduced from Ref. 2
with permission of Wiley-Blackwell.)
Fraction of time
Peak-to-peak value
p-p exceeded
2 × rms (±1σ) 0.32
4 × rms (±2σ) 0.046
6 × rms (±3σ) 0.0027
6.6 × rms 0.001 (1000 ppm)
8 × rms 60 ppm
10 × rms 0.6 ppm
12 × rms 2 × 10–9 ppm
Most noise can be accurately modeled as white. Also, many forms of noise,
such as thermal noise within detectors, can be effectively modeled as Gaussian.
There is no necessary correlation between the white noise model and the
Gaussian noise model. The oscilloscope trace in Fig. 5.5 shows the characteristic
appearance of Gaussian noise.
White noise without a dc component has a zero average over time, and its
peak is infinite (though with zero probability). The preferred way to characterize
this noise is by its rms value denoted by σ, which is also called the standard
deviation. Table 5.3 shows peak-to-peak values of Gaussian noise, along with the
fraction of the time that the peak-to-peak (p-p) value is exceeded.
A good estimate of the rms amplitude of the noise can be obtained by
estimating the difference between the maximum and minimum amplitudes in an
oscilloscope trace and dividing this difference by six: rms amplitude = (peakmax –
peakmin)/6. This can also be done by looking at amplitudes in the data set.
Noise can be expressed as a voltage, a current, or a power, and is best
expressed as a spectral quantity if its magnitude depends upon frequency. The
expression for a mean square noise voltage (equivalent to a noise power) looks
like
T
1
v = (vi − vavg ) = (vi − vavg ) 2 dt .
2 2
(5.12)
T0
Figure 5.5 Gaussian noise oscilloscope trace and accompanying probability distribution.
(Reprinted from Ref. 3 with permission from John Wiley & Sons, Inc.)
Its units are volts squared. In the expression, vi is the instantaneous voltage,
vavg is the average voltage, and T is the observation time. The rms noise
is vn = vn2 . Current may also be used according to Ohm’s law, v 2 = i 2 × R 2 .
If the various noises in a system have different origins, they may be
considered independent or uncorrelated. The noise powers then add algebraically,
while the noise voltages (or currents) add in quadrature:
If the noises are partially correlated, emanating from the same source, a
correlation coefficient must be included in the equation, and the resulting noise
will be greater. In any case, the total noise will not exceed the algebraic sum of
the noise voltages or currents.
∞
1
2
B= G ( f )v( f ) 2 df , (5.14)
G ( f o )vo 0
∞
1
G ( f o ) 0
B= G ( f )df . (5.15)
G(fo) 3-dB
bandwidth
Actual response
Noise
bandwidth
Frequency
relationship differs and must be derived via integration. If the noise is nonwhite,
integration must be performed. ENBs can also be specified in terms of time
constants. The common two-pole low-pass filter with equal time constants
(Bessel filter) has B = 1/(8τ), and the simple single-pole filter has B = 1/(4τ).
Note that the ENB is always stated in frequency space at frequencies below
1012 Hz in the electronic realm. Do not confuse this bandwidth with the passband
of an optical filter operating at frequencies greater than 1012 Hz, whose center
and passband are specified in units of wavelength, nm or μm.
Note that since there is no explicit frequency term, Johnson noise is “white”
at least to 1012 Hz. Integrating over frequency to obtain the mean-square noise
voltage:
v 2j = 4kTRB , (5.17)
where B is the effective noise bandwidth, defined above. The rms Johnson noise
voltage is the square root of the mean square noise voltage, or
v j = 4kTRB . (5.18)
4kTB
i 2j = . (5.19)
R
hf
kT → ( hf / kT )
. (5.20)
e −1
Then,
x
v 2j = 4kTRB ,x
(5.21)
e −1
where x = hf/kT.
It is easy to show that Eq. (5.21) reduces to Eq. (5.17) by applying the
approximation e s ≈ 1 + s when s is small, that is, much less than 1. This
substitution is only necessary for frequencies greater than 1012 Hz, so it is not
significant for most applications. This expression also indicates the linkage
between thermal noise and blackbody radiation.
In most practical applications, the noise bandwidth is established by an RC
circuit time constant. Under these circumstances, the mean-square Johnson noise
is
∞
S ( f ) df
v 2j = , (5.22)
0
1 + (2πfRC ) 2
which integrates to
kT
v 2j = , or
C (5.23)
v j = kT / C .
Note that the expressions in Eq. (5.23) are independent of resistance R, but
dependent upon circuit capacitance C. As an example, if T = 300 K and
C = 1 picofarad (pf), then vj= 64 μV rms. Using the identity q = CV, the noise can
also be expressed as 6.43 × 10–17 coulombs (C), or about 400 electrons.
S s ( f ) = 2qI dc , (5.24)
where q equals the charge on an electron, 1.60217733 × 10–19 C, and Idc equals
the direct current flowing across the potential barrier.
The mean square shot noise current is
2
ishot = 2qI dc B . (5.25)
KI α
S( f ) = dc
, (5.26)
fβ
where K = a “constant” for the particular technology/noise process, 1.25 < α < 4
(usually 2), and 0.8 < β < 3 (usually 1).
Note the spectral frequency dependence of this noise, which has been
demonstrated to occur at frequencies as low as 10–5 Hz and below. This is a
particularly nasty noise, ultimately limiting dc and low-frequency detection and
amplification. This form of fluctuation is ubiquitous, found throughout nature. A
few of the manifestations of 1/f noise are seen in Table 5.4.
The low-frequency form of 1/f noise is often called “drift.” Other names are
“pink noise” (β = 1), “red” or “brown noise” (β = 2), “excess,” “flicker,” and
“contact.” Causes in semiconductor detectors, for example, include nonohmic
contacts and surface impurities.
This noise is particularly insidious, as the noise power is constant in each
frequency decade. The total noise is the integral of the power spectral density
over frequency, which is proportional to ln(fhigh/flow) for pure 1/f noise. Evaluation
is easy for ac-coupled systems, but in direct-coupled systems, we must choose
some number other than zero for flow. Unless a system has a particularly long
integration time, flow of 0.1 Hz is probably sufficient. While integration of
“white” noise over time reduces white-noise effects, integration fails to decrease
1/f noise due to the decrease in flow with increased integration time.
Waves on a beach
Fluctuation in axon membrane
Earthquakes
Economic variables
Ecological time series
Self-organizing systems
Fluctuations in human heart rate
Photon counting
Most music (not Metallica!)
Frequency of rotation of earth (β = 2)
Feedback controls in nuclear reactors
Base arrangement of DNA sequences
Traffic flow (both vehicular & network)
Motion of man standing on one foot
includes variations in the carrier lifetime. The expression for G-R noise can be
complicated, and the specific form depends upon the detector configuration. An
example of the power spectral density of G-R noise for an extrinsic
photoconductor (to be described later in this chapter) is
4 R 2i 2 τ
SG - R ( f ) = , (5.27)
N (1 + 4π2 f 2 τl2 )
where
Note that G-R noise is frequency dependent (nonwhite). Its Bode plot looks
like that of a low-pass filter.
4kT 2 KB
S ΔT 2 ( f ) = , (5.28)
K 2 + 4π 2 f 2 H 2
where
square fluctuations in the photon arrival rate are equal to the arrival rate. The rms
photon noise equals the square root of the number of photons:
(Δn) 2 = n , (5.29)
ex
(Δn) 2 = n . (5.30)
ex − 1
SIGNALmax
LSB = , (5.31)
2n
where
n = number of bits,
SIGNALmax = the full-scale signal (amps, volts, or electrons), and
LSB = the magnitude of the least-significant bit.
Since the quantization noise is proportional to the LSB, the larger the number
of bits, the lower the quantization noise.
Johnson noise, G-R noise, and 1/f noise are usually uncorrelated and
therefore add in quadrature as shown in Fig. 5.7. This figure is typical for a
photoconductive detector. The three frequencies labeled f1, f2, and f3 are the
“corner” frequencies on the curve. In particular, since a plot of responsivity ℜ
versus frequency looks like the G-R noise curve, the region where the D* is
nominally flat extends from f1 to f3 and is the typical choice for operation.
Noise figure NF is the log (base 10) of the noise factor, given by
vn2 + in2 R 2
Tn = . (5.34)
4kRB
Figure 5.8 Noise that is white over three different bandwidths. (Reprinted from Ref. 3 with
permission from John Wiley & Sons, Inc.)
Curve Bandwidth
Upper 200 KHz
Center 20 KHz
Lower 2 KHz
Figure 5.9 1/f noise at three different bandwidths. (Reprinted from Ref. 3 with permission
from John Wiley & Sons, Inc.)
Curve Bandwidth
Upper 2 KHz
Center 200 Hz
Lower 20 Hz
Unlike the case for white noise, limiting bandwidth does not proportionately
reduce peak amplitude for 1/f noise.*
Finally, Fig. 5.10 presents a comparison of sinusoidal signals within white
and 1/f noise, respectively. In both cases, the signal-to-noise ratio is
approximately 1. As seen in the figure, white noise has a “furry” or “fuzzy”
quality in the trace, while 1/f noise is “jumpy.”
Figure 5.10 (a) White noise and (b) 1/f noise for SNR = 1. (Reprinted from Ref. 3 with
permission from John Wiley & Sons, Inc.)
*
From the frequency plane perspective, the smoothing shown is the result of filtering out higher-
frequency terms. From the time-plane perspective, rapidly changing signals are smoothed out by
the energy storage elements in the filter (capacitors and inductors).
k
xi − 0.5
y= , (5.35)
i =1 k / 12
where x = RND(1). Use of this series generates random numbers y with mean = 0
and standard deviation σ = 1. The starting point is k uniformly distributed
numbers xi, where x is between 0 and 1. A value of k = 12 is suggested, which
will give maximum values for y of 6 at the 3σ point. If we wish to generate a new
random variable y′ with mean m and standard deviation σ, we form the
expression
y′ = m + σy . (5.36)
αΦRT
ΔT = Td − To = , (5.37)
1 + ω2 RT2 H 2
where
τT = RT H , (5.38)
αΦRT
ΔT = Td − To = . (5.39)
1 + ω2 τT 2
H
RT
HEAT SINK Ad
To
Td
Figure 5.11 Thermal circuit.
when conduction and convection to the surroundings are minimized. One way to
achieve this goal is to place the detector in a vacuum, support it with low-
conductivity materials, and utilize small thin connecting wires. It can be shown5
that under these circumstances
1
RT = , (5.40)
4ασAd Td 3
K = 4ασAd Td 3 . (5.41)
Under the conditions described, the K in Eq. (5.41) is the only conductance
between the detector and its surroundings. The input power producing a
temperature change ΔT may be expressed as
Φ = K ΔT . (5.42)
Setting the signal-to-noise ratio equal to one so that Φ=NEP, and assuming
the limiting case in which the noise is due only to fluctuations in the incident
power, we obtain:
ΔT = ΔT 2 . (5.43)
The mean square value of the temperature fluctuations in the incident beam
may also be expressed as
4kTd 2
ΔT 2 = B, (5.44)
K
Φ = 4kTd 2 KB . (5.45)
kTd 5 σAd B
NEP = ΔΦ 2 = 4 , (5.46)
α
The maximum D* that can be achieved for a 300 K detector (Ad = 1 cm2,
B = 1) is thus 1.8 × 1010 cm·Hz1/2/W. This is called the Havens limit for a thermal
detector. If the detector element is cooled, this limit increases dramatically,
reaching a D* of nearly 1015 cm·Hz1/2/W at 4 K and close to 1019 cm·Hz1/2/W at
0.1 K. At these low temperatures, other noises predominate; we cannot reach the
theoretical maximum.
ΔV
S= (V/deg), (5.47)
ΔT
METAL 1
J1@T1 I J2@T2
METAL 2
6
Figure 5.12 Thermoelectric circuit. [Reprinted from Optical Radiation Measurement
series, Vol. 4, W. Budde, Physical Detectors of Optical Radiation, p. 101 (1983).]
METAL 1
J1@T1 J2@T2
METAL 3 METAL 3
V
6
Figure 5.13 Open-circuit thermoelectric pair. [Reprinted from Optical Radiation
Measurement series, Vol. 4, W. Budde, Physical Detectors of Optical Radiation, p. 101
(1983).]
1 dQ
Π= , (5.48)
I dt
using multiple thermocouple junctions are called thermopiles and are used in
most thermoelectric transducers.
Some of the theory behind thermoelectric detectors can help to understand
their operation. Going back to the thermal equations common to all detectors, we
have
and
α ΦRT
ΔT = (ac case). (5.50)
1 + ω2 τT 2
and
α SRT
ℜ(ω) = (ac case). (5.52)
1 + ω2 τT 2
R S 2Td
ΔT = αΦRT 1 − T , (5.53)
R
Figure 5.14 Thermopile configurations: (a) Coblentz, (b) Schwartz (c) wirewound, and (d)
evaporated.7 [Reprinted from Semiconductors and Semimetals series, Vol. 5, N. B.
Stevens, “Radiation Thermopiles,” pp. 300–304 (1970).]
Resistance 10 Ω to 2 kΩ 2 to 20 kΩ
Thermopile detectors are extremely versatile due to their small size, low cost,
and wide wavelength range of operation at dc and room temperature. They can be
ruggedized to survive space applications such as horizon sensing and earth
radiation budget measurements. Some terrestrial uses of thermopile detectors are
shown in Table 5.9.
5
5
5
4 SEMI
5
3 T
260 280 300 320 340 36
Figure 5.15 General characteristics of resistance as a function of temperature for metal
and semiconductor materials used in bolometers.
VB RL ΔRB
Vs = ΔV = . (5.55)
( RB + RL ) 2
VB RΔR VB ΔR VB RoβΔT
Vs = ΔV = = = . (5.56)
( R + R) 2 4 R 4 R
Substituting for ΔT from Eq. (5.50) and recognizing that for small ΔT, R~Ro,
the bolometer voltage responsivity ℜv is
Vs VB αβRT
.
ℜv = = (5.57)
Φ 4 1 + ω2 τT 2
dPs
p= C/cm2K, (5.58)
dT
Q = pAd ΔT , (5.59)
where Ad = the sensitive area of the detector and p = the pyroelectric coefficient.
The pyroelectric current ip is the product of radian frequency ω and charge:
i p = ωQ = ωpAd ΔT . (5.60)
6
Figure 5.17 Pyroelectric coefficient versus temperature. [Reprinted from Optical
Radiation Measurement series, Vol. 4, W. Budde, Physical Detectors of Optical Radiation,
p. 129 (1983).]
Figure 5.18 (a) Initial circuit and (b) equivalent circuit of a pyroelectric detector, with the
8
current generator in parallel with a capacitor and load resistor. [Reprinted from
Semiconductors and Semimetals series, Vol. 5, E. Putley, “The Pyroelectric Detector”
(1970).]
i p RL
v= , (5.62)
1 + ω2 RL 2C 2
where RLC = the circuit’s electrical time constant τ and ip is given by Eq. (5.60).
Therefore, the expression for voltage may be rewritten as
ωpAd ΔTRL
v= . (5.63)
1 + ω2 τ2
Substituting for ΔT from Eq. (5.37) and applying the definition of thermal
time constant, the voltage responsivity may be expressed as
αωpAd RL RT
ℜv = . (5.64)
1 + ω2 τ2 1 + ω2 τT 2
In this case, the responsivity increases from zero to a flat region, then
decreases at even higher frequencies. The width of the flat region depends upon
the separation of the thermal and electrical time constants. The generic voltage
responsivity behavior is shown in Fig. 5.19, while Fig. 5.20 depicts voltage
responsivity as a function of electrical frequency with load resistance as the
parameter.
p
FM = , (5.65)
εC ′
where
Figure 5.20 Typical voltage responsivity curve for pyroelectric detectors.6 [Reprinted from
Optical Radiation Measurement series, Vol. 4, W. Budde, Physical Detectors of Optical
Radiation, p. 131 (1983).]
TGS is the acronym for the organic compound triglycine sulfate. It depoles
(loses its internal charge) readily, requiring the periodic or even continuous
application of an electric field to maintain operation. It is also sensitive to
moisture and needs protection. The D* can be enhanced by doping the material
with L-alinine.
SBN is strontium barium niobate, a mixture of the general form
SrxBa1–xNbO3. Polyvinylidene fluoride, PVF2, is a plastic film (tradename Kynar)
which can be cut and formed into custom configurations. Other pyroelectric
materials include lead zirconate titanate (PZT), ceramic, barium titanate, and
barium strontium titanate (BST).
hc 1.2398
λc = = , (5.66)
Eg Eg
Eg = Gap energy
Ep = Photon
energy
Ef = Fermi
level
Valence band
Figure 5.22 Illustration of valence and conduction bands, with Eg the energy necessary to
promote an electron from the former to the latter.
Another way to think about the concept is simply to recall that when Ep =
hc/λ is greater than or equal to Eg, for a given material, electrons generated by
photons at wavelength λ will possess the energy necessary to elevate into the
conduction band.
In Fig. 5.22, the excess energy Ep – Eg appears as heat. In the case of a
photovoltaic cell (to be discussed in greater detail later) this thermalization loss
causes the cell’s voltage and power to decrease. It is one of the two primary loss
mechanisms responsible for the fairly low peak theoretical efficiency of about
28% for simple photovoltaic cells.
A number of interesting intrinsic semiconductor materials are candidates for
optical radiation detection, as shown in Table 5.11. For example, silicon is seen
to require a maximum wavelength of 1.1 μm, 1100 nm, and this leads to the
dramatic falloff on the right side of the photon detector curve seen in Fig. 5.1.
Most of these materials may be categorized in a straightforward manner, but the
last combination deserves special mention. Mercury telluride (HgTe) is classified
as a semimetal, with a small negative energy gap. This means that HgTe is a
conductor at room temperature, albeit a rather poor one. Cadmium telluride
(CdTe) is a semiconductor which has been exploited for visible radiation
detection. When the two tellurides are combined as Hg1–xCdxTe, then the energy
gap depends on x, the fraction of CdTe in the mix. A mixture where x = 2 is
common and yields a detector with response out to about 12 μm. The longest
wavelengths that are practical with this trimetal detector are about 25 μm, and
cold temperatures are required for effective operation.
− Eg / kT
ni 2 = constant × T 3 × e , (5.67)
where
Looking at the two temperatures for InSb in Table 5.12, it is apparent that the
T in the exponent in Eq. (5.67) has more influence on the calculated value of ni
than does the T3 term. Since we want sufficient carriers available for photon
excitation even at low values of ni, we must cool low Eg detectors that are
designed for long-wavelength operation. This fact explains the need, for
example, to cool HgCdTe detectors to temperatures of 77 K or lower. Detectors
should be cooled so that
Eg 600
kT < or T < , (5.68)
25 λc
valence band to acceptor levels, leaving behind a hole for conduction, as shown
schematically in Fig. 5.25. Table 5.13 lists some of the impurity dopants and
their levels and cutoff wavelengths.
The maximum practical doping for extrinsic materials is about 1 ppm, or
about 6 × 1017 atoms/cm3. At room temperature, they are nearly all used up. As
the temperature increases, the material reverts to an intrinsic conductor.
There are a number of other interesting semiconductor materials, among
them PbxSn1–xTe, that have made good photovoltaic detectors. Many alloys can
be formed from combinations of materials in groups III to V or II to VI in the
periodic chart. The materials in Table 5.14 below have been successfully used in
the fabrication of heterostructure alloys for solid-state sources and detectors.
Some are useful in the ultraviolet, others in the visible, and many others in the
infrared.
Table 5.13 Donor and acceptor levels for germanium and silicon.
in Germanium in Silicon
Dopant eV λc (μm) eV λc (μm)
Au 0.15 A 8.3 0.54 A 2.3
Cu 0.041 A 30 0.24 A —
Zn 0.035 A 35 0.26 A —
Hg 0.087 A 14 none —
Cd 0.055 A 22 0.3 A —
Ga 0.011 A 112 0.0723 A 17.8
B — — 0.045 A 27.6
Al — — 0.0685A 18.4
In 0.011 A — 0.155 A 7.4
S — — 0.187 D 6.8
As 0.013 D — 0.054 D 23
Sb 0.0096 D 129 0.039 D 32
Note: “A” denotes an acceptor and “D” denotes a donor.
Table 5.14 Semiconductor materials used for solid-state sources and detectors.
AlP
AlAs
AlSb
GaN
GaP
GaAs
GaSb
InP
InAs
InSb
CONTACT
σe = qμ n n . (5.70)
σe = qμ n ( n + Δn) , (5.71)
Δσe qμ n Δn
= . (5.72)
σe σe
η Eq τ l
Δn = . (5.73)
z
Δσe qμ n ηEq τl
= . (5.74)
σe σe z
Δσe −ΔR
= , (5.75)
σe R
with the negative sign indicating that the relative change in resistance has
opposite slope to the relative change in conductivity.
We wish to derive an expression for the voltage responsivity ℜv at
wavelength λ. To do so, we consider the placement of a photoconductive detector
in a circuit, as shown in Fig. 5.27.
RL
VS = VB , (5.76)
RL + RD
where
VB = bias voltage,
RL = load resistance, and
RD = detector resistance.
The optional capacitor placed in the circuit does not factor into the analysis,
but is included because values of Vs can be very large, and VB can get as high as
200 V. Placing a capacitor in the circuit allows a modulated signal. Note that the
device is symmetrical, and that the polarity of the applied bias in Fig. 5.27 is
unimportant.
We need an expression for ΔVs, the change in output signal voltage due to a
change in resistance. Differentiating Eq. (5.76), we obtain
−VB RL
ΔVs = × ΔRD . (5.77)
( RL + RD ) 2
Note that VB /(RL+RD) is Idc, the dc current flowing through the detector, so that
− I dc RL
ΔVs = × ΔRD . (5.78)
( RL + RD )
and applying the results of Eqs. (5.75) and (5.78) to the definition of voltage
responsivity, we obtain
ΔVs I R R ληqμ n τl
ℜv = = dc L D . (5.79)
Φ hczAd σe ( RL + RD )
ΔVs I dc Rληqμτl
ℜv = = , (5.80)
Φ 2hczAd σe
τl
G= , (5.81)
τtr
l2
τtr = . (5.82)
μVB
λΦ λΦ τl μVB
is = ηq G = ηq ⋅ 2 . (5.83)
hc hc l
λΦ τl μVB R
Vs = ηq 2 , (5.84)
hc l 2
and
λ τ μV R
ℜv = ηq l 2 B . (5.85)
hc l 2
Note that the current responsivity ℜi may be obtained directly from Eq.
(5.83) by dividing the signal current by the power term, such that
ℜI = ηq(λ/hc)G A/W.
We can maximize ℜv by increasing VB, but if Joule heating occurs and the
PC detector heats up, decreasing its resistance, we will burn it out!
λ kT kT
in 2 = 4q ηqΦ G 2 + qG 2 N ′ + + B. (5.86)
hc qRD qRL
The limiting noise is G-R noise from radiation (signal plus background).
When G-R noise overpowers all other noises, the rms signal current is
λ
is 2 = is = ηqΦ G , (5.87)
hc
λ
in 2 = in = 4q 2 ηΦ G 2 B , (5.88)
hc
i Ad B
D* = s ,
in Φ
where Ad is the detector area. Noting that Φ = EAd, and rearranging terms in Eqs.
(5.87) and (5.88), D* for the photodetector becomes
η λ
D *BLIP (λ, f ) = . (5.89)
4 E hc
This quantity is called D*BLIP because, as discussed in Sec. 5.3, the limiting
noise arises from incident photons. Equation (5.89) expresses a quantity
referenced to a particular wavelength and having a specific modulation frequency
f. Expressing this irradiance in terms of photon incidence Eq we also obtain
λ η
D *BLIP (λ, f ) = . (5.90)
2hc Eq
LIQUID
NITROGEN
COLD SHIELD
COLD FILTER
WINDOW
DETECTOR
VACUUM
wavelength can be obtained. The doped silicon detectors are currently in favor
for focal plane applications as on-chip signal processing can be accomplished
with conventional silicon technology. The lead-salt detectors will operate at room
temperature, albeit poorly, but do much better when cooled to –193° C. Most
InSb and HgCdTe detectors prefer 77 K while the extrinsic detectors based on Si
and Ge require even lower temperatures. The lead-salt detectors have somewhat
slower response times than the others.
hc 1239.8
λc = = , (5.91)
φ φ
1
P ( En ) = ( En − E f )/ kT
, (5.92)
1+ e
FERMI-DIRAC FUNCTION
1
100K
400K
0.8
300K 0K
PROBABILITY
200K
0.6
0.4
0.2
0
0 0.5 1 1.5 2
ENERGY
where
10
Table 5.17 Nominal composition and characteristics of various photocathodes.
(Reprinted by permission of Burle Technologies.)
Dark
Wavelength Responsivity Quantum
Nominal emission
of maximum at λmax efficiency
composition at 25° C
response (nm) (mA/W) at λmax (%)
(fA/cm2)
Ag-O-Cs 800 2.3 0.36 900
Ag-O-Rb 420 1.8 0.55 —
Cs3Sb 330 64 24 0.3
Cs3Sb 400 42 13 0.2
Cs3Sb 340 50 18 0.3
Cs3Bi 365 2.3 0.77 0.13
Ag-Bi-O-Cs 450 20 5.6 70
Cs3Sb 440 48 14 4
Cs3Sb 480 20 5.3 —
Cs3Sb 440 48 14 3
Cs3Sb 440 23 6.7 —
Cs3Sb 490 83 21 1.2
Na2KSb 420 64 19 0.0003
K2CsSb 400 95 29 0.02
Rb-Cs-Sb 450 92 25 1
Na2KSb:Cs 420 64 19 0.4
Na2KSb:Cs 420 64 19 0.3
Na2KSb:Cs 420 44 13 —
Na2KSb:Cs 530 44 10.3 2.1
Na2KSb:Cs 575 37 8 0.2
GaAs:Cs-0 850 119 17 92
GaAsP:Cs-0 450 61 17 0.01
In.06Ga.94As:Cs-0 400 50 15.5 220
In.12Ga.88As:Cs-0 400 69 21 40
In.18Ga.82As:Cs-0 400 42 13 75
Cs2Te 250 25 12.4 0.0006
CSI 120 24 20 —
Cul 150 13 10.7 —
K-Cs-Rb-Sb 440 84 24 —
These conversion factors are the ratio of the radiant responsivity at the peak
of the spectral response characteristic in amperes per watt (A/W) to the luminous
responsivity in amperes per lumen (A/lm) for a tungsten lamp operated at a color
temperature of 2856 K.
A newer class of photoemitters known as negative electron affinity (NEA)
materials feature a special surface treatment on a p-type semiconductor substrate
to “bend” the band structure. In extreme cases, the vacuum level is below the
bottom of the conduction band. The advantages of this NEA photocathode
include longer wavelength operation and higher quantum efficiency. Figure 5.32
schematically depicts photoemission from these materials.
J = CT 2 e( −φ / kT ) , (5.93)
where
RICHARDSON EQUATION
1E-6
1E-8
DARK CURRENT (A)
1E-10
1E-12
1E-14
1E-16
1E-18
1E-20
1E-22
200 220 240 260 280 300 320 340 360
TEMPERATURE (K)
λ
is = ηqΦ q = ηqΦ . (5.94)
hc
If signal current flows through load resistor RL, then the signal voltage is
λ
vs = RL ηqΦ . (5.95)
hc
Applying Eq. (5.95) and the results from Eqs. (5.18) and (5.25), the noise
voltage is
1/2
λ 4kT
vn = RL 2qid + 2q 2 ηΦ + B . (5.96)
hc RL
λ
η qΦ
SNR = hc . (5.97)
1/ 2
2 λ 4kT
2qid + 2q ηΦ + B
hc RL
The ultimate limit is achieved when the dark current shot noise and the
Johnson noise from the load resistor can be reduced, leaving only the signal-
dependent shot noise. Under these conditions, the SNR is
ηλΦ
SNR = . (5.98)
2hcB
δ( n +1)
NF = . (5.99)
δ n (δ − 1)
δ
NF = . (5.100)
(δ − 1)
This noise factor is quite small, typically less than 1.2. The gain of the
electron multiplier is essentially noise free. Table 5.18 lists some of the positive
and not-so-positive characteristics of photomultiplier tubes.
Photomultiplier tubes have found a number of different uses in areas
including photon counting, spectroradiometry, and imaging. In the latter, many
PMT-based devices have been replaced with solid-state imagers.
Figure 5.34 Photomultiplier dynode arrangements: (a) circular-cage type, (b) box-and-grid
type, (c) linear-focused type, (d) venetian blind type, (e) fine mesh type, and (f)
microchannel plate.11 (Reprinted by permission of Hamamatsu Photonics K. K.)
11
Figure 5.35 Typical spectral responses of common photocathode materials. (Reprinted
by permission of Hamamatsu Photonics K. K.)
(1) Free electrons in the n region are attracted to the positive charge in the p
region and drift over.
(2) Free holes in the p region are attracted to the negative charge in the n
region and they drift over.
(3) Carrier drift leaves the n region with a net positive charge and the p
region with a net negative charge. The crystal stays neutral with no net
carrier gain or loss.
In the n-type material (dopants are As, Sb, and P), the electrons are the
majority carriers and the holes are the minority carriers. In the p-type material
(dopants are Al, B, In, and Ga), the holes are the majority carriers, and the
electrons are the minority carriers. Majority carriers are far more mobile than
minority carriers, and they are the primary contributors to current flow.
The barrier height depends upon the donor and acceptor levels and
concentrations. This is shown schematically in Fig. 5.36. The region between the
n and p regions is called the depletion region, and there is an electric field across
it.
The barrier height is calculated as
kT nn p p
φ≈ ln 2 , (5.101)
q ni
where
bias source repels carriers from the other side of the junction (n-type) and vice
versa. The consequence is a low current flow due to conduction by minority
carriers. Increasing the barrier height widens the depletion region.
The equation expressing the I-V characteristic of a p-n junction diode is
derived from a continuity equation:
βqVkT
I d = I o e − 1 , (5.102)
where
q = electronic charge,
k = Boltzmann’s constant,
T = absolute temperature in degrees kelvin,
V = applied voltage,
β = a “constant” to make the equation fit the data, sometimes called the
“ideality” factor, and
Io = reverse saturation current.
At 300K, q/kT is equal to 38.7. The “constant” β varies with applied voltage.
It is typically 1, but can be as high as 3.
Curves for the I-V equation are shown in Fig. 5.39 for various values of Io. A
large Io yields a large reverse current and a small forward voltage drop and vice
versa.
Figure 5.39 Current-voltage curves for a p-n junction with various Io.
n p Dn pn D p
Io = q + Ad , (5.103)
Ln Lp
where
q = electronic charge,
np = minority carrier (electrons) concentration in the p-region,
Dn= Einstein diffusion constant for electrons,
Ln = minority carrier (electrons) diffusion length in the p-region,
pn = minority carrier (holes) concentration in the n-region,
Dp = Einstein diffusion constant for holes,
Lp = minority carrier (holes) diffusion length in the n-region, and
Ad = detector area.
kT
D= μ, (5.104)
q
and has units of cm2/s with μ being carrier mobility. Like D, it may be
subscripted with n or p to specify electrons or holes. The minority carrier
diffusion length is
L = D τl , (5.105)
REVERSE SATURATION CURRENT vs. TEMP
0.0001
1E-05
1E-06
1E-07
1E-08
1E-09
260 270 280 290 300 310 320 330 340 350 360 370 380 390 400
TEMPERATURE
Figure 5.40 Reverse saturation current versus temperature for a typical p-n junction.
qV
I = I o e βkT − 1 − I g , (5.106)
λ
I g = η qΦ q = η q Φ. (5.107)
hc
β kT I o + I g
Voc = ln . (5.108)
q Io
If Ig >> Io, which is almost always the case, then Voc is logarithmic with
radiant power as seen in Fig. 5.42. If Ig << Io, then Voc becomes linear with
incident power as shown in Fig. 5.43. If Io ~ Ig, operation is intermediate between
linear and logarithmic.
Photovoltaic detectors are also commonly used as power generators. Placing
a load resistor RL directly across the detector causes the I-V curve to enclose an
area, as seen in Fig. 5.44. Short-circuit current, graphically depicted as the
location where the I-V curve crosses the current axis (V = 0), and open-circuit
voltage, where the I-V curve crosses the voltage axis (I = 0), form the two
extreme points of the (inverted) I-V curve; in between, both current and voltage
are available simultaneously. This is the requisite condition for power generation,
and devices operating in this region are commonly called solar cells. The lower
curve in Fig. 5.44 is the power versus voltage characteristic for a particular cell,
in which the maximum power is achieved at about 0.48 V.
The open circuit voltage Voc is a result of the forward-biased rolloff seen
generically in Fig. 5.39. For silicon, this voltage is close to 0.6 V as in Fig. 5.44.
The short-circuit current Ig is proportional to the incident radiation Φ as shown in
Eq. (5.107), and the radiation is in turn proportional to the area of the cell. The
cell in Fig. 5.44 with an Ig of 50 mA is quite small with an area of about 1 cm2.
Cells for typical commercial solar panels or modules have an area of around 100
or more cm2, producing a short-circuit current of around 5 A. A typical panel has
perhaps 72 cells connected in series to produce an open-circuit voltage near 40 V
and a short-circuit current near 5 A. The ratio of the maximum power to the
product VocIg is called the “fill factor,” (not to be confused with fill factor in
CCDs, below) which is typically on the order of 0.75 or so.
50
40
POWER x 2
MILLIAMPS
30
20
10
0
0 0.1 0.2 0.3 0.4 0.5 0.6
VOLTS
Figure 5.44 Photovoltaic solar cell operation (lower curve) and I-V characteristic (upper
curve).
0.8
RESPONSIVITY (A/W)
0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1
WAVELENGTH (um)
Figure 5.46 Ideal (straight line) and actual (curved line) current responsivity versus
wavelength.
(1) shot noise from current flow across the potential barrier,
(2) Johnson noise, and
(3) 1/f noise.
The most important noise source is shot noise. Shot noise current arises from the
signal, background, and the device’s own dark current. The rms shot noise
current is
4kTB
iJ = . (5.110)
Rdyn
The dynamic resistance is low at forward bias and high at reverse bias. It is
calculated as
dV βkT − qV /( Ad kT )
Rdyn = = e . (5.111)
dI qIo
dV βkT
Rdyn = = . (5.112)
dI qI o
1/2
( constant )( I dc df )
i1 f = . (5.113)
f
λ
is = ηq Φ. (5.114)
hc
1/ 2
λ
ishot = 2ηq 2 ΦB . (5.116)
hc
Inserting the expression for the signal current and the shot-noise current Eq.
(5.116) (for the infrared) in Eq. (5.8) for D* results in
* ηλ
DBLIP (λ , f ) = , (5.117)
2 Eback hc
* λ η
DBLIP (λ , f ) = . (5.118)
hc 2 Eq ,back
Comparing Eqs. (5.89) and (5.117), and (5.90) and (5.118), we see that the
photovoltaic detector is better (higher D*) than the photoconductive detector by a
factor of 2. The physical reason for the difference is the absence of G-R noise
in the photovoltaic detector. Recombination takes place within the
photoconductive detector itself; in the photovoltaic detector, it occurs in an
external ohmic circuit where carriers are not statistically correlated. An exception
to this rule arises when photoconductive detectors are operated in “sweepout”
mode. In this case, carriers are swept out of the detector before they recombine,
and the D* is equal to that of a photovoltaic detector.
Configuration Comments
p-i-n Built-in depletion region, with an intrinsic layer of Si
between p and n materials
Avalanche Operate under a large reverse bias; fabricated from Si, Ge,
and InGaAs. Can be cooled to 77 K and biased beyond
breakdown point.
Schottky barrier Created by depositing a thin semitransparent metal electrode
on top of a semiconductor material. Particularly useful for
large-area UV detectors.
Inversion layer Created by doping top layer on p-type silicon with a material
having a positive charge to form an n-type material.
Quantum efficiency approaches 1 at short wavelengths.
Ultraviolet Include Schottky barrier, front-illuminated PtSi, and
AlxGax–1N
Infrared Include InGaAs, InAs, HgCdTe, and PbSnTe. Require
cooling beyond 3 μm.
Position sensing Output as a function of position on the detector, often used in
photodiodes (PSPD) tracking applications.
Material Comments
Monocrystalline silicon n/a
Poly- or multicrystalline silicon n/a
5.7.1 Introduction
Many applications require spectral and spatial information that the use of a
single-element detector cannot reasonably provide. In order to obtain this
information, some imaging systems use single detectors along with scanning
optics and other components. More commonly, linear or area (2D) arrays of
detectors are used, with the area array in widest use. A comparison of the
functional differences of single and multiple detectors is shown in Table 5.23.
Array imagers are relatively recent attempts to emulate human vision. Other
innovations throughout history are listed in Table 5.24.
This section will review concepts important to array detection, including
history, basic array parameters, device architecture, and specific array types
along with their applications. Due to the large and growing amount of material on
this subject, the reader is strongly encouraged to investigate the references at the
end of the chapter.
5.7.2.1 History
One of the earliest man-made optical detectors was photographic film. The
treatment here is brief because film has been supplanted in most scientific and
consumer applications by solid-state electronic detectors, as described in this
chapter. It is still used, however, in some specialized applications. For more
information on photographic film, the reader is referred to two thorough books
on the application of photography in science and engineering (Ray, 1999) and the
theory and technology of photography (Stroebel et al, 2000).
D = − log(T ) . (5.119)
A plot of D for each step versus the logarithm of the corresponding exposure
log(E) is known as the “D-logE” or “H-D” (for Hurter Driffield) curve. The H-D
curve essentially shows output (optical density) versus input (log exposure)
calibration. It depends not only on film type, of course, but also on processing
conditions (chemical temperatures, processing time), exposure spectral
characteristics, density measurement procedures, and many of the variables
associated with photography. The main characteristics of the H-D curve are a
saturation toe at low exposure (the “base + fog” level, typically a D of about 0.1),
a linear segment for moderate exposure, and a saturation shoulder at high
exposure (typically a D of about 3.0). The equation of the linear segment is
D = γ log( E ) + D0 , (5.120)
where D0 is the projected intercept (usually negative) on the density axis (not the
“base + fog” level), and gamma γ is the gain of the film, synonymous with its
contrast; i.e. a low gamma means a low-contrast film. The H-D curve describes
“macrocalibration” for relatively large-area measurements of a millimeter or
more. If calibration of film is attempted for “micro” conditions of measurement
of a tenth of a millimeter or less, other factors come into play, for example, the
adjacency effect where nearby exposure affects the density at a point of interest.
5.7.2.6 Summary
Film remains a unique image recording mechanism in that the recorded and
developed image is the archival medium itself, i.e. no additional processing is
necessary to save the original image for long periods of time. It is also an
efficient detector in terms of its combination of large format and high resolution.
For example, a large-format aerial photograph 10 × 10 in. with a resolution of 10
μm contains some 645,160,000 resolution cells, or pixels. At the time of this
writing, such large monolithic electronic detector arrays are not possible
(although mosaics of individual arrays can achieve this size).
However, the inconvenience, delay, and cost of chemical processing and
conversion to a digital format by scanning and the associated quantitative
difficulties in image measurements have seriously disadvantaged film relative to
5.7.3.1 History
One of the earliest array devices was the linear photodiode array marketed by
EG&G Reticon. It used silicon photodiodes in the photon flux integrating mode
coupled with field-effect transistor (FET) switches. These were driven by
clocked shift registers and stored electrons in the device capacitance until
readout. During the readout phase, electrons were transferred to an output signal
amplifier. Similar linear-array devices today find application in spectroscopy,
astronomy, grocery and department store scanners, and many other products.
The first large-area imaging arrays were built in the late 1960s using x-y
addressable photodiodes, phototransistors, and photoconductors. They were not
particularly successful due to responsivity nonuniformities and spatial noise. In
addition, only instantaneous readouts were employed, and the signal could not be
integrated over time. Advantages were good area utilization and random access
to pixels within the array.
Charge-coupled devices (CCDs) allowed integration of photosensitive
elements with complex readout mechanisms. Photodetectors within a CCD
operate in integration mode, with the outputs serially clocked at high rates to a
single readout circuit. CCDs having on the order of 100 megapixels (MP) have
been fabricated; this number will increase with technology development over
time. Some of the parameters characterizing imaging detector arrays are listed in
Table 5.25.
Serial clocks
SERIAL
REGISTER
Analog
Output
FULL
Parallel ARRAY
clocks
Parallel
shift direction
Figure 5.47 Full-frame readout.
upward to the serial-shift register in time with the parallel clock signals depicted
at left. Upon reaching the serial-shift register, charge is read serially into the
output amplifier. The process is repeated until all rows are read out. It requires
some form of shuttering of the array, so that light collection does not occur
simultaneously with charge transfer.
Frame transfer architecture, depicted schematically in Fig. 5.48, is similar to
the full-frame architecture with the inclusion of a storage array. As in the case of
the full frame, rows of charge are shifted toward the serial register, where charge
from each element is read out sequentially in accord with serial clock signals.
The storage array seen in Fig. 5.48 is not photosensitive, and light collection and
integration in the image array can occur simultaneously with integration in the
storage array. As a result, there is no need for a shuttering scheme.
A weakness of this mechanism is that integration within the image array is
still occurring while transfer to the storage array takes place, resulting in image
smear. Because twice the silicon is used as in the full-frame device, a frame-
transfer device costs more. On the other hand, higher data rates are achieved.
In the interline transfer architecture, shown schematically in Fig. 5.49,
photosensitive areas and readout registers are arranged in successive columns.
After detection and integration takes place in the photosensitive area, charge is
transferred to column registers, which are then clocked to the serial register as in
the previous two architectural techniques. This architecture allows very fast
response time, but is comparatively difficult to fabricate. Although image smear
as in the frame transfer device is not entirely eliminated, the amount of smear is
reduced.
Serial clocks
SERIAL
REGISTER
Analog
STORAGE Output
Parallel ARRAY
clocks (MASKED)
(storage)
Parallel
clocks IMAGE
(image) ARRAY
Parallel
shift direction
Figure 5.48 Frame-transfer readout.
CID. Pixels are individually addressable in a CID. The charge remains intact in
each pixel after its signal level has been determined, making for a nondestructive
readout. Row and column electrodes are shifted to ground to make way for
collection of the next image frame, and the charge is “injected” to the substrate.
This capability for individual pixel control is valuable in many imaging
applications. For instance, long exposures to low-light-level sources can allow
optimum exposure of a particular target, such as in astronomical use. If other
objects of interest within the sensor’s field of view appear during the long
integration, the pixels containing such images can be read out while the low-
light-level target continues to integrate on the array. At brighter target levels,
blooming is less likely to occur than in the CCD, due to the fact that charge
overload is confined to a single pixel.
CMOS. Like CIDs, CMOS devices are x-y addressable, allowing individual pixel
information to be read out without reading out the entire array. A disadvantage of
CMOS is high readout noise unless an amplifier is part of the configuration of
each pixel (that is, unless the device is an active-pixel sensor) which, as stated
above, reduces the device fill factor. Additional signal processing circuitry
providing functions such as thresholding, edge detection, and motion detection
can be incorporated into the CMOS chip, allowing CMOS to compete on a
performance level with scientific CCDs.
In addition to the common monolithic CMOS structure (detector and readout
together on one chip) CMOS devices are available in hybrid form (two chips
bump-bonded together). The latter structure allows a 100% fill factor in the
visible spectrum.
5.7.3.4 Comparison
When presented with a choice of a camera based on one of these three
technologies—CCD, CID, or CMOS—one must consider one’s application
relative to the technology and its cost. As in most radiometric system and
measurement problems, there are tradeoffs to be made. While not purporting to
offer the final word on technology selection, Table 5.26 presents a comparison of
some of the characteristics of these technologies. This comparison is by necessity
general; many specific tradeoffs can be made to enhance performance and reduce
cost, depending upon application.
best. The situation is changing, however, as digital cameras are being developed
in competition with film cameras for photogrammetric purposes. Because of the
application, these digital cameras would have to offer better radiometric
sensitivity than that obtained using digitized film images and would provide the
capability for self calibration.13
Alternatively, three separate CCDs may be used to provide color imaging,
with filters designed to pass radiation in the red, green, and blue portions of the
spectrum. This approach involves more hardware than the one discussed above,
but will become more cost effective as the price of this technology comes down.
utilize aluminum gallium nitride (AlGaN) as the detector material due to its
sensitivity in the UV and insensitivity at longer wavelengths. Flame detection,
astronomy, and undersea communications are some applications that benefit from
these devices. Current technical challenges include tailoring the cutoff
wavelengths of AlGaN p-i-n photodiodes by controlling alloy composition,
allowing AlxGa1–xN arrays to cut off between 227 nm and 365 nm.
5.7.8 Summary
The above descriptions of array detectors, their uses, and some comparisons are
only an overview. Trading off parameters for a specific application is often
necessary, and there is no particular “right” answer to many detector selection
problems. A thermographer examining a housing structure for sources of heat
leakage can purchase an uncooled thermal instrument for far less cost than a
cooled device and yet receive adequate imagery for that task. An astronomer
observing a distant galaxy or low-light-level astronomical source may find a
cooled photodetector array to be his only reasonable option in the infrared, but
might he choose a CID over a CCD for observations in the visible?
Questions such as these are important for the student or professional wishing
to become adept at systems analysis and systems engineering, two of the most
important disciplines making use of the art of radiometry.
A number of books have one or more chapters dealing with optical radiation
detectors. These include the following:
R. W. Boyd, Radiometry and the Detection of Optical Radiation, John Wiley &
Sons, New York (1983).
A. Chappell, Optoelectronics: Theory and Practice, McGraw-Hill, New York
(1978).
R. D. Hudson, Infrared System Engineering, John Wiley & Sons, New York
(1969).
D. Malacara, Physical Optics and Light Measurements, Academic Press, New
York (1988).
Several major handbooks offer articles and/or chapters dealing with optical
radiation detectors. These include the following:
J. S. Accetta, and D. L. Shumaker, Eds., Infrared and Electro-Optical Systems
Handbook, Vol. 3, SPIE Press, Bellingham, Washington (1993).
M. Bass, Ed., Handbook of Optics Vol. 1, McGraw-Hill, New York (1995). Parts
5, 6, and 7 contain Chapters 15 through 25, all pertinent to detectors and
detection.
W. L. Wolfe, Ed., Handbook of Military Infrared Technology, Chapters 11 and
12, Office of Naval Research, Washington D.C. (1965).
W. L. Wolfe and G. Zissis, Eds., The Infrared Handbook, Chapters 11–16, ERIM
and SPIE Press (1978).
Another rich source for detector information, and probably the best for
assessment of the current state of the art, is documented in the Proceedings of the
various conferences of SPIE. SPIE sponsors several major conventions per year,
each having one or more conferences on detectors. Compilations of the best
Proceedings papers, along with other critical papers, are gathered in SPIE’s
Milestones series http://spie.org/x649.xml.
References
6.1 Introduction
Radiometric instruments vary in what they are intended to measure, how they do
it, how complicated and expensive they are, how rugged, and in a number of
other ways. In this chapter, the simplest of radiometers is considered, the
components of radiometers are described, and spectral radiometers are covered.
6.3.1 Introduction
Chapter 2 introduced basic optical concepts relevant to radiometry as well as
some simple radiometer configurations. They will be reviewed here in
preparation for more detailed discussion of instrumentation. The main issues of
radiometer optics are the placing of stops and pupils, possible reimaging, use of a
field lens, and the reduction of scattered light.
In this case the detector is the aperture stop, and the hole is the field stop.
Since there are no lenses, there are no pupils. The flux on the detector from a
point source is given by
I s Ad
Φd = , (6.1)
d2
Ls Aap Ad
Φd = , (6.2)
d2
where Ls is the radiance of the source and Aap is the area of the aperture.
The field of view is “fuzzy”: that is, different parts of the detector view
somewhat different parts of the field of view because the detector is the field
stop. Note that if you make the detector a little smaller, the field of view is a little
smaller, so the outer portion of the detector views a different portion of the field
than the inner portions.
A reasonable improvement is obtained by the use of two apertures, as shown
in Fig. 6.4. The detector is no longer the field stop. For a point source, and A2
smaller than the active area of the detector, A2 is the aperture stop, A1 is the field
stop, and the flux on the detector is given by:
I s A2
Φd = , (6.3)
d2
Ls A1 A2
Φd = , (6.4)
d2
where d is the distance between the two apertures. The field is still a little fuzzy.
One example of such a radiometer that uses a single detector and two
apertures and baffles but no lenses, is a pyrheliometer (fire + sun, measuring the
sun’s fire) used by the World Meteorological Organization and shown
schematically in Fig. 6.5. The angles made with the detector by A, B, and C
should be about 1, 2, and 4 deg, respectively.
Lens
Detector
arrangement, the field is well (not fuzzily) defined, but the detector still “sees”
radiation coming at different angles from different parts of the field of view. This
is the case as long as the detector is the field stop. The advantages are that this is
a simple system—one lens, one detector, and associated electronics. The
disadvantage is that the detector averages over different parts of the field of view.
In this system the irradiance on the entrance pupil (the lens) from a point
source is given by
I s τatm
Eo = , (6.5)
d2
where Is is the source intensity, τatm is the atmospheric transmittance and d is the
distance from the source to the entrance pupil. The flux on the entrance pupil is
just the lens area Alens times this, and the flux on the detector is just the lens
transmittance τlens, times that:
If the source is extended, the distance from the instrument to the source is
irrelevant; it is the focal length that counts.
Aperture
stop
Field
lens
Detector
Φ d = τo Ls Ad Ω fsd , (6.8)
Afs
Ω fsd = , (6.9)
d2
where d is the distance from the detector to the field stop. Thus, one can also
write the power on the detector as
τo Ls Ad Afs
Φd = . (6.10)
d2
τo Ls Ao A fs
Φd = . (6.11)
f2
If the image of the objective exactly fills the detector, either of Eqs. (6.10) or
(6.11) is accurate.
Detector
Mirror
Eye
Detector
Reflective
chopper
Reference
Mirror
Eye
Figure 6.9 Viewing with a chopper.
6.3.10 Choppers
Most good radiometers use choppers, or radiation modulators. Of course there
are advantages and disadvantages. The ac signal from a chopped radiometer
provides a discriminant against a static background. It allows the use of drift-free
ac amplifiers. It avoids the low-frequency part of the 1/f noise region. It provides
the ability to use a synchronous detector, and it provides an a-b type of
comparison measurement. On the other hand, choppers reduce the available flux
by a little over 50%; they can be noisy both electrically and acoustically, they can
have reliability problems, and there can be phase noise if things are not exactly
right.
Most choppers are just spinning blades, driven by an electric motor. In this
form they do not use much energy. There are many types available on the market.
Some can be resonant devices that oscillate in and out of the beam, perhaps
tuning forks. These, too, are readily available from companies listed on the
Internet. Browse for “optical choppers”; both types are advertised.
The final step is to analyze the system with one of several programs that can
calculate the amount of radiation on the detector from out-of-field sources as a
function of their angle out of field. And there is no substitute for a measurement
if the application warrants it. A number of stray-light-analysis systems are listed
on the Internet. Search for “stray light optical techniques.” The reduction of stray
light is an iterative technique that is shortened by experience. The analysis
programs tell you how you are doing.
The Lyot stop is a special kind of stop useful in controlling scatter. It was
used by Lyot to measure the solar corona. He imaged an interior disk inside the
system to the front. The image of the disk matched the solar disk, allowing him
to see the light outside the main disk of the sun—the corona and flares. The
image of the disk became the entrance pupil, an image of the aperture stop. This
is one application of using an entrance pupil to control stray light.
6.3.12 Summing up
The design and use of radiometric instruments is simple in principle, but difficult
in practice. It is a science of precision and care. One of the maxims is to think of
everything. One way to do this is to write the radiometric equation for the
radiometer:
This means that the output voltage (or other electrical signal) is a function of
wavelength, time, spatial coordinates, angular coordinates, the phase of the
moon, the relative humidity, the temperature (of the radiometer and the source
and the background), and everything else you can think of. Then test the
radiometer against all of these variables. The devil is in the details—and so are
good results!
6.4.1 Introduction
Spectral instruments include both those that make relative measurements and
those that make absolute measurements. The first type measures how much
radiation there is in each part of the spectrum, but only on a relative basis. They
are useful for chemical analysis, for instance. The absolute instruments, generally
called spectroradiometers, measure the spectrum, but also give information about
how many watts (or equivalent radiometric quantity) are at each wavelength.
They are calibrated and more difficult to use. This section is about various types
of spectroradiometers. Another way to view them is that they are radiometers
with several, or many, narrow spectral bands. The different types vary from each
other according to the way they obtain the spectrum. That can be with prisms,
To get the optimum response from a spectrometer, one should make the
width of the entrance and exit slits the same. One should also just fill the entrance
slit with the foreoptics. Underfilling robs you of input; overfilling gives no
advantage and even increases stray light.
where m is the order number, an integer, λ is the wavelength of the light, and
θi and θd are the angles of incidence and diffraction, respectively. For a given
wavelength and angle of incidence there are many maxima at various diffraction
angles that correspond to different values for m and θd. The 0th order includes all
wavelengths and is just regular reflection (or transmission). For a given incidence
angle, the first order, m = 1 for a wavelength λ, has the same diffraction angle as
the second order and half the wavelength. This is overlapping of orders and limits
the free spectral range.
The resolving power Q is given by
λ
Q= = mN , (6.14)
dλ
where m is the order number and N is the number of grating lines illuminated.
The throughput is the slit area times the projected area of the used portion of
the grating divided by the focal length of the optics:
The free spectral range of the prism is unlimited, while that of the grating is
limited by multiple orders, the different maxima corresponding to different
values of m. Gratings can be transmissive, reflective, and even concave to
incorporate some of the focusing properties of the system. The resolving power
of a grating is approximately constant across the spectrum, whereas the prism’s
Q varies. The resolving power of a prism is given by
λ bdn
Q= = , (6.16)
dλ dλ
where b is the length of the base of the prism (actually the length of the beam
near the base) and dn/dλ is the change in refractive index of the prism with
respect to wavelength.
The throughput is dictated by the slit area and the focal ratio of the focusing
optics:
Mirror systems have more variability, but generally there is an entrance slit,
an exit slit, one or two mirrors, and a grating. The Czerny-Turner, Fig. 6.15, was
developed by Marianus Czerny and Francis Turner, a onetime student of
Czerny’s and later a researcher at Eastman Kodak and professor at the College of
Optical Sciences, The University of Arizona. There are two versions; one might
be called straight and the other crossed, as in Fig. 6.16. The crossed version uses
the two paraboloids closer to on axis and should therefore have smaller
aberrations. Of course, if aberrations are not a problem, the paraboloids can be
replaced by spheres. In that case, and depending upon the speed of the optics,
spherical aberration rather than coma will be dominant. It is shown with a
reflective grating, but it is not hard to imagine how this setup can be used with a
transmissive grating. The essence of the design is that both the collimating optic
and the focusing optic are off-axis paraboloids. However, they are off axis in
opposite directions so that the comatic aberrations offset each other.
The Fastie Ebert mount is shown in Fig. 6.17. It has the advantage of using
only one mirror, although the mirror has to be larger than either of the mirrors of
the Czerny-Turner system. Again, the mirror can be either a sphere or a
paraboloid, and again the comatic aberrations tend to offset each other.
Obviously, this cannot be used with a transmissive grating.
6.4.4 Spectrometers
These devices come in several different varieties. They can be single or double
pass through the dispersers, like the Wadsworth (single) or the Littrow (double).
They can be single or double beam, as shown below. The double beams were
developed to get an automatic referencing system. In some older single-beam
systems, one had to wrap a cord around the screw that rotated the prism in just
the right way. This was to “program” the slit to account for the variation in
intensity from the source. One also had to run a calibration run, and then a
sample run, and manually do the ratioing. The double-beam system eliminates all
this. There are many ingenious variations on the several simple systems shown
here. One beam passes through the sample, whatever it may be. The other is the
reference beam and is just in air in the spectrometer. The output of the sample
beam is divided by the output of the reference beam to give the transmission of
the sample. There are several ways to do this, but the best way is to use the same
detector for both beams, perhaps with an appropriate chopper. This avoids the
obvious problem of detector matching, initially and repeatedly. The generic basic
two-beam instrument is shown in Fig. 6.19. The output coming from the exit slit
of the monochromator is collimated and divided into two beams by a divider. It
can be a (semitransparent) beamsplitter or a bladed chopper. The one beam goes
straight through; the other is diverted by two mirrors and then combined with the
first beam by a beam combiner just like the beam divider—a synchronized
chopper or a plane-parallel plate. If the divider and combiner are choppers, the
electronics can just take the ratio of signals. If beamsplitters are used, some
technique must be used to “tag” each of the beams. The beams need not be
collimated, but could be relayed by lenses for compactness.
Other commercial instruments can be found on the internet by entering the
search term “spectrometers.” Many will show up with diagrams, prices, and
advertising!
There is one subtlety about double-beam systems that I (Wolfe) encountered
quite by accident. I was participating in a military study devoted to the detection
of land mines. Someone suggested that soil was transparent around 4 μm. I
scoffed, but I went home to prove my point. I took one inch of certified backyard
dirt into our double-beam spectrometer. Lo and behold, there was a transmission
peak at 4.3 μm! This could not be! So I took one inch of aluminum plate and
made the same measurement. Same result. I pondered this for a while. Clearly
aluminum is not transparent at 4.3 μm. The answer was in the reference beam.
The atmosphere is very absorbent at 4.3 μm due to carbon dioxide. What I was
measuring was the absorption in the reference beam that translated to apparent
transmission in the overall measurement. I have since seen this phenomenon in
other double-beam measurements. If there is extra transmission at 4.3 μm in the
scan of an optical material you have ordered, be skeptical. This illustrates at least
one principle of radiometry: do not simply believe whatever you measure; make
sure it makes sense. By the way, this problem can be obviated by filling the
reference tube with nitrogen.
opposite directions, thereby combining the light and eliminating or canceling the
spread of the spectrum.
6.4.6 Arrays
Monochromators can also be used with detector arrays. The arrangement
includes the foreoptics described above, an entrance slit, collimating optics, and
the array, which functions as a set of exit slits. If the array is linear, the system
can be viewed simply as a monochromator with many exit slits, each one sensing
a different wavelength. This system obviously has the multiplex advantage, but
does not have the throughput advantage. The detector elements can usually be
sized correctly for the proper operation, but apertures can also be used with a
concomitant loss of signal.
6.4.8 Filters
Filters come in a variety of types. They can be based on absorption, interference,
and even scattering. Absorption and interference filters are the main candidates in
spectroradiometry. Almost any kind of bandpass can be generated by a proper
thin-film design. They can be narrowband, broadband, angle tolerant, multiple
bandpass, etc. Then they can be put in a filter wheel to obtain a spectrum of a
sort, although not a continuous spectrum. They have good throughput but do not
have as good resolving power as prisms and gratings. They can come in
segmented wheels or circular or linear variable filters. These latter devices are
interference filters with layer thickness variation around the circumference or
along the length. They therefore have a spectral band that varies with either angle
or position.
Their characteristics vary by design. The circumference of a circular variable
filter (CVF) is given by
Δλ
C= Do , (6.18)
dλ
where Δλ is the spectral range, dλ is the resolution and Do is the diameter of the
aperture stop, where it is placed. The diameter of the CVF is this value divided
by π. A representative filter has the following characteristics: the spectral range
is 2.5 to 14.1 μm, Q is 67, FWHM is 1.5%, and resolution varies from 0.03 μm at
2.5 μm to 0.2 μm at 14.1 μm. The CVF is divided into three segments that cover
from 2.5 to 4.3, 4.3 to 7.7, and 7.7 to 14.1 μm. It can be operated at the focal
plane, as the manufacturer recommends, but it should be operated at a field stop
rather than the focal plane. Since it would then be in converging light, the
resolution will be affected. Of course other varieties are available, semicircles for
instance. Typically the spectral range is an octave, the resolving power is about
50, and the diameters are about 4 cm. Specific devices can be found on the
Internet by searching for “optical filters” or “variable optical filters.” If you leave
out “optical,” all sorts of electrical filters will appear in the search results.
Optical filters are optically and mechanically simple, cheap, rugged, and easy
to automate, but they have poor stray-light suppression and limited resolving
power.
6.4.9 Interferometers
The main interferometers used for spectral analysis are the Michelson and the
Fabry-Perot, or as some insist, the Perot-Fabry. The former is a two-beam
instrument, the latter a multiple-beam device. The multiple-beam system has
greater resolving power, but smaller free spectral range.
λ σ σ 5000
Q= = = = . (6.19)
d λ d σ 2δ λδ
6.4.11 Fabry-Perot
This interferometer is generally used for high-resolution spectroscopy. It is
essentially a pair of plane-parallel plates, as shown in Fig. 6.21. The light from
the source is collimated by the first lens, and passes through the first plate after a
little lost reflection. Light reaches the second plate and is reflected back to the
first, which reflects it back to the second, which reflects it back to the first. This
interferometer does have high resolution because of the multiple-beam
interference. The basic equation for the transmission of the Fabry-Perot is
τo
τ(λ) = , (6.20)
4ρ 2
1+ sin φ
(1 − ρ) 2
and
(φ1 + φ2 )
φ = 2πnd σ cos θi − , (6.21)
2
σ λ ρ
Q= = = mπ , (6.22)
dσ dλ 1− ρ
T = τo AΩ , (6.23)
1
Δσ = . (6.24)
2d
The Fabry-Perot has a relatively poor throughput because of the requirement for
collimation between the plates. But it is better than a prism or grating.
Final note: Acousto-optical tunable filter (AOTF) devices for spectroscopy of all
kinds are relatively new. “Classical” spectrometers employing prisms and
gratings have seen many improvements. There are pros and cons for both. They
are a tool that should be in the radiometrist’s kit and are described further in
Wolfe and in Chang, below.
References
241
a voltmeter with only 1% accuracy is far less desirable than one with 99%
accuracy. The systematic error in a measurement is closely related to calibration
of the apparatus used to conduct the measurement.
Random (type A) errors are those that vary in an unpredictable manner when
the same quantity is repeatedly measured under identical conditions. They are
revealed only by multiple measurements. Precision is a term often associated
with random errors; a measurement is considered precise if it is repeatable. We
may employ several methodologies to reduce random errors, enhancing the
measurement precision. First, we can use a finer scale division, i.e., more bits, to
reduce granularity in the measurement. Second, we can reduce some of the
inherent noise in the measurement process by filtering, cooling, shielding, etc.
The most important tool to reduce random errors is statistical analysis. We take
multiple readings and perform analysis to reduce the effects of “noise” and
increase our confidence in the measurement.
To understand the total uncertainty in a measurement, we must consider both
the systematic and the random error components. Figure 7.2 shows a set of
“measurements” with various combinations of accuracy and precision. The
average (x and y) of the third pattern lies very close to the center of the target.
Errors can be specified as absolute, the magnitude of the error in the appropriate
engineering units, or as a relative or fractional error, usually in percent.
1
Measurement uncertainty = . (7.1)
SNR
Tool Formula
Mean
x i
x = i
, N = number of
N
measurements
( x − x)
2
i
Variance (sample)
σ 2= i
N −1
2
σ= σ
Standard deviation
σ
Standard deviation of mean σx =
N
1 2
σ 2y ( τo ) =
N −1
( yk +1 − yk ) . (7.2)
2( N − 1) k =1
1
( )
2
σ 2( N ) =
N
yk − y . (7.3)
N −1 k =1
SIGNALmax
LSB = , (7.4)
2n
where LSB is the least significant bit and SIGNALmax is the full-scale reading.
If this quantization error is significant in a measurement, the use of more bits
(larger n) is indicated. If the presence of sharp quantization levels is apparent in
your data and you cannot get more bits, you can artificially smooth the data by
adding noise. This is called dithering, used to “smear” and mask the appearance
of discrete levels.
One is often tempted to reject data that appears to be “out of line.” The
general recommendation is DON’T. The data has warts, leave them alone. If you
really must, there are reasonable guidelines for data rejection. Don’t make the
mistake of expunging data from your data set, just set the outliers aside. After all,
there might be something really interesting there. The most popular tool is the
Chauvenet criteria for rejection of outliers. Here we reject a data point if the
probability of observation of the suspect point is less than ½N where N is number
of observations. Then we can recompute mean and standard deviation. The
criteria are shown in Table 7.4 above.
Application of Chauvenet’s criteria affects standard deviation more than
mean. It is interesting to test the dataset with and without data rejection to see if
it really matters! Usually it doesn’t. A final note: even if tempted, do not apply
this technique more than once. If you do, the data will probably be skewed. If one
keeps rejecting data, eventually one will have only a single data point remaining.
To minimize systematic errors, calibrate frequently with suitable apparatus,
as noted above. Table 7.5 lists several sources of systematic error.
The evaluation of systematic error is, in practice, a judgment call, based upon
an investigator’s familiarity with previous data, instrument behavior,
manufacturers’ specifications, handbook reference data, and calibration. Current
recommended practice is to estimate systematic errors at the 1σ (68%
confidence) level.
Systematic errors in optical radiation measurements tend to be large because
the quantities involved are a function of everything in the world; the following
factors are the most common:
(1) Wavelength (broadband or monochromatic)
(2) Power or energy level
(3) Linearity
There are two secondary error categories: illegitimate errors and model
validity. Illegitimate errors include blunders, mistakes, computational errors
(roundoff, etc.), and chaotic errors. Model validity has to do with our
preconceived ideas about the results of a measurement, which influences the way
we conduct the measurement. While these errors may be significant or even
overwhelming in a particular situation, none are essential contributors to the
limits of radiometry and can be eliminated with proper care.
When all of the various error sources have been identified, they must be
combined into an overall uncertainty estimate. Errors propagate according to
well-known formulae as shown in the following tables. Table 7.6 presents the
formulas for single-valued functions.
For combinations of functions, the worst-case ultraconservative treatment is
to directly add the errors. This method would be correct if all of the errors were
correlated. Table 7.7 presents the formulas for functions in which independent
variables are combined.
σq σx σ y σz
= + +
q x y z
Multiplication and division (q = xy/z)
σq σx
Power (q = xn) =n
q x
σc = σ 2r + σ2s . (7.5)
The 1σ level implies a confidence level that is only 68%. To give more
confidence in the measurements, we multiply sigma by a coverage factor k. A k
of 2 is 2σ, where σ is the standard deviation of the measurement. Use a suggested
confidence factor k of 2, which gives a confidence of approximately 95%.
A recommended format for reporting results shows the measured value with
the standard (1σ) uncertainty and a coverage factor k as follows:
In standards work, authors frequently split the random and systematic uncertainty
components and report both. For further details, consult Taylor.2
2 2 2
σq σ σy σ
Multiplication and division (q = xy/z) = x + + z
q x y z
which is not particularly useful. We must integrate over area, solid angle, and
wavelength to get the integral form:
These equations are valid only for incoherent radiation; another layer of
complexity is added when speckle and interference effects are added.
There are many alternate forms of the measurement equation. For a single
detector with spectral responsivity ℜ(λ), it simplifies to:
where Φλ(λ) = spectral radiant power incident on the detector and ℜ(λ) = the
radiometer spectral power responsivity.
For a detector with spectral responsivity ℜ(λ) and a narrow-band source or
filter:
where τ(λ) = spectral transmittance of the filter and Δλ = bandwidth of the source
or filter.
For a laser line where the spectral linewidth is so small that the responsivity
of the detector/radiometer can be considered constant, then
SIGNAL = ℜΦ . (7.11)
1
SIGNAL =
d2 ℜ ( λ ) I λ (λ) dAd λ , (7.12)
where
Many other forms can be developed depending upon the chosen measurement
configuration.
Similar to the measurement equation is the range equation, which gives the
distance at which a source will generate a specified SNR. It is useful to visualize
which system parameters are important in particular applications. A general form
is provided by Hudson:3
1 1
1 π Do τo 2
1 1 2
R = ( Is τa ) 2
( D *) 2
(7.13)
4 ( f #) Ω SNR B
TERM 1 2 3 4
This range equation can take many specific forms to determine SNR or range
for various source/radiometer spatial and temporal configurations. Note, also,
that the term intensity implies a point source.
There are several alternate means of assessing the performance of a
radiometric system based on what are called “noise equivalent” quantities, whose
input values produce an SNR of 1 (signal equals noise). Noise-equivalent power
(NEP), for example, was defined in Chapter 5 as
in i
NEP = = Φ n ;or
ℜ is
v v
NEP = n = Φ n ;or (7.14)
ℜ vs
Ad B
NEP = .
D*
It is usually applied to detectors, but may also be useful for systems designed
to measure radiant power.
We can also define a noise-equivalent irradiance or flux density (NEI or
NEFD), which is frequently used to characterize systems for detection of distant
small sources:
E v NEP
NEI = =E n = . (7.15)
SNR vs Ad
Note that the term NEI does not reflect our symbol for irradiance E.
For an unresolved target that underfills the FOV, Hudson3 gives the noise-
equivalent irradiance (W/m2) as
4( f #)Ω1/ 2 B1/2
NEI = , (7.16)
πDo D * τo
4( f #) B
NETD = λ2
, (7.17)
∂M (λ, TB )
Ad τoptics (λ) D * (λ ) d λ
λ1
∂T
The last of these five definitions is the most useful, as it adds reality to the
measurements that we conduct. The common thread in the definitions of
calibration is the involvement of a standard. A valid measurement is inextricably
linked to the calibration process and therefore to physical standards. How well
we can measure is closely related to the quality of the standards we employ, and
NIST does not define nor enforce traceability except in its NVLAP
laboratory accreditation program. Moreover, NIST is not legally
required to comply with traceability requirements of other federal
agencies; nor do we determine what must be done to comply with
another party’s contract or regulation calling for such traceability.
7.6.1 Introduction
The selection of an appropriate configuration to conduct a calibration is governed
both by the desired measurements and by the availability of appropriate facilities
and standards. There are a number of different ways to set up a calibration or
comparison source and a radiometer. The five basic calibration configurations
that follow are the most common combinations of radiometer aperture-stop and
field-stop considerations and source distance and size. One of them should
suffice for any radiometric instrumentation calibration. If at all possible, use two
or more configurations to gain additional confidence in the calibration.
f2
ℜ E = ( SIGNAL) , (7.19)
I
where f = focal length of the collimator (m) and I = source intensity (W/sr) = LAs.
The result is again irradiance responsivity ℜE in SIGNAL/(W/m2). The
advantages to this approach include: (1) a controllable atmosphere (vacuum
chamber, if necessary), (2) a controllable background, (3) the distance S is not in
the equation, (4) you can use almost any small source, and (5) the radiometric
signals tend to be larger. The disadvantages include (1) the need to know the
focal length f (typically a one-time measurement), and (2) the collimator/chamber
hardware can get very expensive. Examples include laboratory calibrators and
low-background test chambers. As to collimators, their basic types are refractive
and reflective.
The basics of a refractive collimator are shown in Fig. 7.6. Its advantages
are: (1) relatively simple alignment due to unfolded path (in the visible, only),
and (2) a simple setup, particularly if off-the-shelf components are used.
The disadvantages include: (1) the wavelength range is limited by the τ(λ) of
the lens material, (2) reflection losses occur due to refractive index, (3) ghost
images arise due to reflections from optical elements, (4) antireflection coatings
are wavelength dependent, (5) chromatic aberration from refractive optical
elements occurs, and (6) difficulties in alignment occur if the lens is not
transmissive. However, most of these disadvantages can be dismissed if
operating at a wavelength for which the collimator components are optimized.
SIGNAL
ℜL = , (7.20)
L
IMAGE
FIELD
STOP
FIELD ANGLE
SOURCE APERTURE
STOP
Figure 7.9 DES calibration configuration (adapted from Wolfe and Zissis).6
The advantages of this configuration include: (1) the distance between the
source and the radiometer is not important, and (2) there is no background due to
the fact that the source overfills the field of view. Disadvantages include: (1) an
intervening atmosphere, and (2) need for a large uniform source. Source
examples include White Sands, New Mexico, and a lake of known surface
temperature for remote-sensing applications, and a large integrating sphere,
blackbody radiation simulator, or a white diffuse panel in the laboratory.
SOURCE
FIELD
STOP
FIELD
ANGLE
APERTURE
STOP
Figure 7.10 NES calibration configuration (adapted from Wolfe and Zissis).6
Figure 7.11 NSS (Jones) calibration configuration (adapted from Wolfe and Zissis).6
SIGNAL A
ℜL = , (7.21)
Ls As
INCOMING BEAM
CIRCULAR
FIELD
STOP
LENS AND
APERTURE
STOP
To minimize effects from laser drift and noise, use a beam-power stabilizer
(expensive) or a beamsplitter and another stable detector to characterize the
beam-power fluctuations during the measurement. In addition, you must ensure
that saturation of either detector does not occur.
7.6.7 Conclusion
The above calibration configurations yield different types of responsivities, but
under many circumstances we can use the simple equation for transfer of radiant
power in an optical system, Eq. (2.47):
Φ = LAΩ .
ℜE ℜL
ℜΦ = = . (7.22)
A AΩ
1 m diameter
T = 2000 K
200 km
Ad B 1cm 2 1 Hz
NEP = = = 1 × 10−10 W . (7.23)
D* 1010 cmHz1/ 2 /W
(2) Does the source represent a point source for this configuration?
The first thing to figure out is if we are dealing with a point source or an
extended source, as we do not wish to use system performance equations
indiscriminately. We know that the source is “small,” given its distance to the
sensor, but we need to determine its relationship to our detector’s size.
Electrical bandwidth B 1 Hz
Because the diameter of the (diffraction-limited) source image is less than the
dimension of our detector, 1 cm or 0.01 m, this source qualifies as a point source.
NEP 1 × 10−10 W
NEI = = = 1 × 10−10 W/cm 2 . (7.25)
Ad 1cm 2
(4) How does this value compare to the NEI obtained from Eq. (7.16),
above?
Ad 1cm 2 1
Ω= 2
= 2
= = 2.78 × 10−4 sr . (7.27)
f (60 cm) 3600
NEI = 10 1/2 −1
= 3.183 × 10−13 W/cm 2 . (7.28)
π(20 cm)1 × 10 cmHz W
*
Note, although this is not a text on system design, the size of the diffraction-limited “blur” plays a
role in sizing a system’s detector(s). Further information on system design is obtained in
references, some listed below and others in the appendices.
This result differs from that in Eq. (7.25) by three orders of magnitude.
Rechecking calculations, we find no error causing a discrepancy this large; what
could the problem be?
To answer, look again at Eq. (7.25). NEI was calculated according to the area
of the detector—not the area of the system entrance pupil Ao. If we repeat that
calculation using the correct surface, we obtain
NEP 1 × 10−10 W
NEI = = = 3.185 × 10−13 W/cm 2 , (7.29)
Ao 314 cm 2
which is much better. Conclusion: Make sure you are addressing problems with
the mathematical expression which corresponds to the setup/configuration you
are analyzing.
The answer to this question requires an inversion of the range equation, Eq.
(7.13). It is:
I s τa πDo τo D *
SNR = . (7.30)
4 ( f #) ΩR2 B
It also requires that we know source intensity within the particular spectral band.
This is obtained through
I s = Ls Ap , (7.31)
Substituting into Eq. (7.30), assuming unity transmittances and a 1-Hz electrical
bandwidth,
I s .....
SNR = .
R 2 .....
I Ωo
E= ,
R2
where Ωo may be thought of as the “unit solid angle,” having value 1 sr.
Applying this notion to Eq. (7.32) allows for dimensional consistency.
In conclusion, while real-world problems may be different from those
described in this section—for example, they may include source/target
parameters that change with time and sources that differ from black/graybodies—
these equations are basic, yet powerful enough to provide the engineer or analyst
a starting point from which to develop solutions.
References
magnification, 54, 57
Kirchhoff, 4 majority carriers, 187
law, 77, 98, 99, 149, 314 Marcel Golay, 236
KTC reset noise, 143 marginal ray, 24
matte reflectors, 58
maximum power transfer theorem,
156
Lambert, 4 of electrical engineering, 172
Lambert-Bouguer-Beer law, 70