This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
(S?2
'—1 , + S?i)
(4.36)
57
4.C. The Fresnel equations
where \{ n2 - k2 - s i
(4.37)
(4.38) At # = 0 the reflectivities are R ±(0) = /?„(()) =[(w - 1 ) 2 + k2]/[(n + I) + k2], which is the same as (4.14). As # increases, the component of the reflectivity perpendicular to the plane of propagation increases monotonically to 1 at # = TT/2. The parallel component first decreases and then increases to 1 at # = 7r/2; however, it never becomes zero, as it does at the Brewster angle when k = 0. The amplitudes of the reflectivities are plotted versus # in Figure 4.7 for a medium with a complex refractive index. In two special cases equations (4.35)-(4.38) can be simplified somewhat. When k «c 1, 2
n2 - k2 - n2,
- (n 2 -sin 2 ^) 1 / 2 ,
- nk/(n2 -sin 2 #) 1 / 2 . (4.39)
At the opposite extreme, when n2 - k2 » 1, (4.40) Figure 4.7. Fresnel reflection coefficients for m = 1.50 + i 1.0 versus the angle of incidence. m = 1.5+M.O 0.8
0.6
0.4
0.2 Parallel
30
60 6 (degrees)
90
58
4. Specular reflection
and, after a little algebra, the reflectivities become
(,cntf
+r
(n+cosftf+k2
11
(n + l/cos&f+k2
V
'
When k # 0 and the incident light is unpolarized, the different phase shifts for the two reflection coefficients cause the reflected light to be partially elliptically polarized. 4.D. The Kramers-Kronig reflectivity relations At ft = 0 the complex amplitude reflectivities for the two directions of polarization [(4.31) and (4.32)] are equal: f(0) = r± (0) = f|((0) = (n -1 + ik)/(n + 1 + ik) s Vei0,
(4.43)
where 77 is the amplitude of f(0), and 0 is its phase. Because n and k are both functions of frequency v, r\ and 0 are also functions of v. This may be made explicit by writing (4.43) in the form r(O,v) = v(v)ei0{v\ or ln[f(O,i/)] = ln[7j(i/)] + /0(i/). (4.44) The reflectivity must obey causality. Thus, as discussed in Chapter 3, the expression for the amplitude reflectivity possesses crossing symmetry; that is, r(0, - v) = f*(0,*0. Hence, according to the discussion in Appendix B, r)(v) and 0(v) must obey relations of the form of (B.10): o v — v 0(,,) = _ _
i-'v » dv'.
(4.45a) (4.45b)
Equations (4.45) are the Kramers-Kronig reflectivity relations. They can be used to calculate the complex dielectric constant from measurements of the normal reflectivity as a function of frequency, ri(v) = y/R(0,v). The normal reflectivity is measured over as wide a range of frequencies as practical. A suitable theory, such as the Drude model (Chapter 3), is then used to extrapolate the measurements to v = 0 and v = oo? and these values are inserted into (4.45b) to calculate 0(v). From r)(v) and 0(i/) the components of the refractive index can
4.E. Absorption bands in reflectivity
59
be found by solving (4.43): « = (1-TJ 2)/(1-2TJCOS0+T72),
(4.46a)
and fc = 277sin
0.15. From (2.69) it may be seen that this requires that the absorption coefficient be so large that the irradiance propagating through the medium is reduced by a factor of e over a distance of less than half a wavelength. 4.E. Absorption bands in reflectivity located in incremental area dA a distance ( away on the white screen at position (x,y,z) = (/*sin0,O, / = ( / 1 - / 2 ) / 2 , and alternately add and subtract equations (8.36) to obtain y V^J
The spectral reflectivity at normal incidence of a material having an absorption band given by a classical Lorentz oscillator with the same parameters as in Figures 3.1 and 3.2 is shown in Figure 4.8. The material is assumed to be in air, so that m = m2:== n + ik. Note that the peak of the reflectivity occurs at a higher frequency than the band center v0 and that the band is broadened. The reason the frequency of the band center shifts is the anomalous dispersion of the real part of the refractive index, which, as can be seen in Figure 3.2, decreases with frequency throughout the main region of the absorption band. The frequency at which n = 1 is known as the Christiansen frequency, denoted by *>c, and the corresponding wavelength is the Christiansen wavelength Ac. If the material has only one absorption band, and is in vacuum, then Figure 3.2 shows that n = l only at vc = vQ, the band center. However, there are two reasons why vc may differ from v0 by a considerable amount in nature. First, most materials have several absorption bands at different frequencies, so that the total refractive index consists of a sum of several terms, each of which is described by a Lorentz oscillator function. In this case the frequency at which the total n goes through 1 will be displaced to the high-frequency side of v0. A second reason is that the material may be embedded in a medium whose refractive index is considerably different from 1.
60
4. Specular reflection
The latter case provides a method for making a narrow-bandwidth transmission filter. Small crystals of suitable material are embedded in a second medium whose refractive index is chosen to make n = 1 at the desired frequency, which also must be away from the center of any strong absorption bands in either material. Rays of light whose frequencies are close to the Christiansen frequency will pass undeviated through the surfaces between the two materials, but rays of other frequencies will be bent out of the beam by refraction. The peak in reflectivity near the center of a strong absorption band provides another method of obtaining a nearly monochromatic beam of radiation without using a prism or diffraction grating. Light from a source giving off a broad range of wavelengths is arranged to be reflected many times from the polished surfaces of an appropriate material, as shown in Figure 4.9. After several reflections the beam consists almost entirely of light of a narrow range of wavelengths around the reflectivity maximum. Before long-wavelength diffraction gratings were widely available, this technique was commonly used to isolate wavelengths in the infrared. The narrow bands of wavelengths produced in this manner are known as restrahlen, a German word literally translated as "residual rays." For this reason the strong Figure 4.8. Spectrum of the Fresnel reflection coefficient at normal incidence for a material having the Lorentz dielectric constant and refractive index shown in Figures 3.1 and 3.2. Note that the reflectivity peak is much broader than, and is displaced to the high-frequency side of, the dielectric-constant peak. 1
0.8
0.6
\
0.4
0.2
•
V :
J
A
0.5
1 v/v
:
1.5
4.F. Criterion for optical
flatness
61
vibrational bands of solids in the infrared are often called restrahlen bands. 4.F. Criterion for optical flatness Even highly polished surfaces have myriads of small scratches and other imperfections, and all surfaces are rough on the scale of the size of atoms, a few angstroms. Thus, an important question arises: Under what circumstances are the equations of this chapter valid? That is, when may a surface be considered smooth? This question may be answered with the aid of Figure 4.10, which shows schematically a wave incident at angle tfona surface whose vertical deviations from a perfect plane are characterized by a height &. The portions of the wave front indicated by the two rays have been reflected from parts of the surface differing in elevation by &. As Figure 4.9. Isolating the frequencies in a restrahlen band by multiple reflections.
Figure 4.10. Schematic diagram of two parts of a wave reflected from a rough surface.
hcose
62
4. Specular reflection
indicted in the figure, the path difference over which the two rays have traveled is AL = 2 & cos #, so that the portions of the wave front they represent differ in phase by A0= 27rAL/A = 4'7r^cos^/A. If A0
5 Single-particle scattering: perfect spheres
5.A. Introduction
A fundamental interaction of electromagnetic radiation with a particulate medium is scattering by individual particles, and many of the properties of the light diffusely reflected from a particulate surface can be understood, at least qualitatively, in terms of single-particle scattering. This chapter considers scattering by a sphere. Although perfectly spherical particles are rarely encountered in the laboratory and never in planetary soils, they are found in nature in clouds composed of liquid droplets. For this reason alone, spheres are worth discussing. Even more important, however, is the fact that a sphere is the simplest three-dimensional object whose interaction with a plane electromagnetic wave can be calculated by exact solution of Maxwell's equations. Therefore, in developing various approximate methods for handling scattering by nonuniform, nonspherical particles, the insights afforded by uniform spheres are invaluable. In the first part of this chapter some of the quantities in general use in treatments of diffuse scattering are defined. Next, the theory of scattering by a spherical particle is described qualitatively, and conclusions from the theory are discussed in detail. Finally, an analytic approximation to the scattering efficiency that is valid when the radius is large compared with the wavelength is derived. 63
64
5. Single-particle scattering: perfect spheres 5.B. Concepts and definitions 5.B.I. Radiance
In a radiation field where the light is uncollimated, the amount of power at position r crossing unit area perpendicular to the direction of propagation ft, traveling into unit solid angle about ft, is called the radiance and will be denoted by /(r, ft). Radiance is often also called specific intensity, or simply intensity, or brightness.
Note the difference between irradiance /, which refers to power per unit area of a collimated beam, and radiance /, which is the uncollimated power per unit area per unit solid angle (Figure 5.1). The power per unit area traveling into a small range of solid angles Aft about some direction ft is /(r, ft)Aft. The irradiance may be considered to be the limit of a highly collimated radiance, /(r,ft o ) = limAri_>o[/(r,fto)Aft] = /(r,ft o )5(ft-ft o ), where 8(x) is the delta function. The delta function, which was introduced by P. A. M. Dirac, is an extremely useful mathematical tool. It has the following properties: 8(x) = 0 everywhere except at x = 0, where 5(0) is infinite in such a manner that its integral over an interval ab containing x = 0 is faS(x) dx = 1. Then for any function fix), f*f(x)8(x) dx = f(0)f*8(x)dx = f(0) if ab contains x = 0, but if not, the integral has the value zero. Thus, if a radiation field consisting of a collimated irradiance /(ft 0 ) is examined by a directional detector with a solid angle of admittance Aft, the response will be proportional to / An /5(ft - fto)dft, which is equal to / if Aft includes the direction from the source ft0, and is zero otherwise. In this chapter a number of particle scattering parameters will be defined for the case where the particles are spherically symmetric. The definitions will be extended to nonspherical particles in Chapter Figure 5.1. Irradiance and radiance.
Irradiance = power/unit area
Radiance = power/unit area/unit solid angle
5.2?. Concepts and definitions
65
6, and to assemblages of irregular, randomly oriented particles in Chapter 7. The geometry of the single-particle scattering problem is shown schematically in Figure 5.2. In the discussions and definitions in this chapter the radiation incident on the particle will be referred to as if it is an irradiance / ; however, it may also be considered to be an increment of radiance /AH. This incident radiation interacts with a particle whose refractive index relative to the surrounding medium is m = n + ik. When an electromagnetic wave is incident on a particle, a certain amount of power is removed, or extinguished, from the incident wave and either is absorbed or has its direction of propagation altered. In this book it will be assumed that the energy removed from the beam can be only scattered or absorbed; phenomena such as fluorescence and Raman scattering will be ignored. However, the energy may be reemitted at a different wavelength as thermal radiation; this topic will be treated in Chapter 13. The scattered radiance is the power per unit area deflected into unit solid angle traveling radially away from the particle and measured at a distance large compared with both the particle size and wavelength. 5.B.2. Cross sections Let the total amount of power of the incident irradiance / that is affected by the particle be PE. Then the extinction cross section is crE = PE/J. (5.1a) A certain portion Ps of PE is scattered into all directions, and the Figure 5.2. Scattering by a single particle. The plane containing / and / is the scattering plane.
66
5. Single-particle scattering: perfect spheres
remainder PA is absorbed by the particle. The scattering cross section is
(5-2a)
Qs =
(5.2b)
QA = °A/
(5-2C)
where
5.B.4. Particle single-scattering albedo and espat Junction The ratio of the total amount of power scattered to the total power removed from the wave is the particle single-scattering albedo, w. From the definitions of the cross sections and efficiencies, w = PS/PE =
(5.3b)
In general, the efficiencies and the single-scattering albedo are functions of wavelength. 5.B.5. Scattering and phase angles Consider the radiance scattered by a particle into some direction of interest that makes an angle 0 with the direction of propagation of the
5.B. Concepts and definitions
67
incident irradiance; 0 is the scattering angle, and the plane containing the incident and scattered propagation vectors of interest is the scattering plane. The complement of the scattering angle is called the phase angle, g = ir — 0. The term comes from the lunar phase angle, which is the angle subtended by the earth and the sun as seen from the moon. The phase angle is the angle between the source and the detector of radiation as measured from the center of the particle. In the literature on scattering by a single particle, the scattering angle 0 is usually used. However, in discussions of diffuse scattering by particulate media, including laboratory and remote-sensing applications, the phase angle g is often more convenient. The phase and scattering angles are shown in Figure 5.2. 5.B.6. Particle phase Junction Suppose the power described by irradiance /(r,ft 0 ) traveling into a direction n o is scattered by a particle at position r into a radiance pattern /(r,ft). The particle phase function pig) describes the angular pattern into which the power Ps = JaQs is scattered. Let (d!P 5 /£/ft)(fl 0 ,n) be the power scattered by the particle from direction ft0 into unit solid angle centered about direction ft, and let g be the phase angle between ft0 and ft. Then p(g) is defined by (5-4) The particle phase function may be described equivalently by p(0). Because the area perpendicular to the incident radiation affected by the particle is crQE, the scattered radiance, or power per unit area scattered into unit solid angle, is
If the particle scatters isotropically, pig) = 1. Because the total power scattered in all directions must equal P s, there is a normalization condition on p(g) such that
f
J
(5.6a)
4
Because the particle is spherically symmetric, the scattered power is independent of azimuth; hence, the integration may be carried out
68
5. Single-particle scattering: perfect spheres
over azimuth, so that d£l = 2TT sin gdg, or d£l = 2TT sin 6 dd, as appropriate, and the normalization condition becomes i Cp(g)smgdg = ^ [Vp(0)sinOde = 1.
(5.6b)
In some references the factor 4TT is omitted in (5.4), in which case pig) is normalized so that fp(g)dfl = 1. The particle phase function is sometimes also known by the unwieldy name of single-particle angular-scattering function.
5.B.7. Asymmetry factors The particle angular-scattering function may be characterized by various parameters, usually referred to as asymmetry factors. One such parameter is the cosine asymmetry factor £, which is the average value of the cosine of the scattering angle d = ir — g:
= \f CQ&0p(0)d£l\/\f p
p(O)sindd0
1 r71' = -
(5.7a)
A similar parameter is the hemispherical asymmetry factor j3. This parameter is usually used in conjunction with the two-stream technique for solving the equation of radiative transfer (see Chapters 7 and 8), in which it is assumed that a uniform radiance is incident on the particle from a complete hemisphere. Then /3 is defined such that the fraction of incident power scattered backward into the hemisphere from which the radiance arrives is (1 — j3)/2, and that scattered forward into the hemisphere into which the radiance is moving is (l + )3)/2. That is, ~2~
=
2TT ./back hemisphere
Jfprward hemisphere
~J^~ ^ b a c k
If p(g) is expressed in terms of cos g instead of g (which can always be done using Legendre polynomials - see Appendix C), it can readily
5.C. Scattering by a perfect sphere: Mie theory
69
be shown that this equation becomes 1+6
/"w/2 /"w/2 rTr/2 rTr/2
[
[ J
& = 0 L=-TT/2JA
r77
f
=0
DfcOS 8)£} PK cosLdLdAcos&dti, A4 7 r
(5.7b)
where cos g = cos(# + A)cos L. A positive value of f or j8 implies that most of the light is scattered into the forward hemisphere, and a negative value means that the particle is predominantly backscattering. For example, if p(g) = l + btcos g, then £ = - bx / 3 , and j8 = - b1 / 4 . If the particle scatters light isotropically, £ = /3 = 0. Note, however, that the inverse may not be true: £ = /? = 0 implies only that the particle scatters symmetrically about the g = TT/2 plane, but p(g) is not necessarily isotropic.
5.C. Scattering by a perfect, uniform sphere: Mie theory
The simplest case of single-particle scattering is that of a plane electromagnetic wave scattered by a uniform spherical particle. This problem was solved by Gustav Mie, and the result has come to be known as Mie theory, although other individuals, especially Debye and Lorenz, may have legitimate claim to having been the first to obtain a solution. Frequently in the remote-sensing literature the term Mie theory or Mie scattering is used to refer to scattering by a particle large compared with the wavelength and of any shape. However, this is an incorrect and corrupt usage of the term, which refers only to scattering by an isolated spherical particle of any size. Mie theory is treated in greater or lesser detail in a number of books, including the work of Born and Wolf (1980), Stratton (1941), and Van de Hulst (1957). The most readable, especially for the novice, are the books by Bohren and Huffman (1983) and Kerker (1969). The Bohren and Huffman book contains a FORTRAN computer program for numerical calculation of the intensity and polarization of light scattered from a sphere. Several of the figures in this chapter are based on calculations that used this program. The solution to the Mie problem is lengthy and complicated, and its details are not particularly instructive nor insightful. Hence, rather than force the reader to, in the words of Bohren and Huffman, "acquire virtue'through suffering," I will simply outline the mathematical procedure, referring the interested reader to the excellent
70
5. Single-particle scattering: perfect spheres
treatments in the books mentioned earlier. Following this summary, the properties of the solution will be discussed in some detail. The same procedure is followed as in the derivation of the Fresnel equations in Chapter 4: First the wave equation is solved in a coordinate system appropriate to the problem, and then the electric and magnetic fields are required to satisfy continuity conditions on the surface of the sphere. The resulting expressions give the extinction and scattering efficiencies and the particle phase function. Because the particle is a sphere, a spherical coordinate system is chosen whose origin is at the center of the particle and whose polar axis is parallel to the direction of the incident radiation. The general solution to the wave equation in spherical coordinates is derived in Appendix C and is
=£ E
S =0/71 = 0
xk/^Jft^
(5.8)
where F(>%#,r/f,f) can refer to either the electric or magnetic field, ^ = 2 T T / A , A is the wavelength, A* is the radial coordinate, ft is the polar coordinate, if/ is the azimuthal coordinate, h{}\3?r) and hffKJTr) are the spherical Hankel functions of the first and second kind, '¥#£(#, ^) and ^* } (#,i/0 are the even and odd spherical harmonic functions, respectively, I is an integer (I > 0), m is an integer (0 < m< l \ and the J / ' S , &% g"s, and 3f9s are constants. As discussed in Appendix C, equation (5.8) describes a series of spherical waves propagating radially inward and outward with respect to the particle. At large distances from the sphere the amplitude of the wave is proportional to 1 / A*, which means that the power falls off as 1 / A*2. The strength of the wave varies with direction. In order to complete the solution, the incident plane wave must also be expressed in a form similar to (5.8) as an infinite series of spherical wave functions. This can be done because the spherical Hankel and harmonic functions form complete, orthogonal sets (see Appendix C). The incident and scattered fields must satisfy various conditions, including that the fields be finite everywhere, that the only field traveling inward far from the particle be the incident irradiance, and that the usual conditions of continuity on the transverse and perpen-
5.D. Properties of the Mie solution
71
dicular components of the electric and magnetic fields (Chapter 4) at the surface of the particle be satisfied. The resulting expressions for the extinction, scattering, and absorption efficiencies and the particle phase function consist of slowly converging infinite series. Except when the particle is small compared with the wavelength, a very large number of terms are necessary to evaluate the expressions, and a computer is required for the solution. Because these expressions do not yield particularly useful insights, they will not be written down explicitly here. Again, the interested reader is referred to the references cited earlier. 5.D. Properties of the Mie solution 5.D.I. General properties
The Mie solutions depend on only two parameters: the refractive index of the particle relative to the surrounding medium, m = n + ik, and the ratio of the particle size to the wavelength. The latter is usually expressed in the form Z=27Tfl/A = 7rZ)/A, (5.9) where a is the particle radius, D = 2a is the diameter, and A is the wavelength in the surrounding medium; X is called the size parameter. The solutions have different properties depending on whether X is small, comparable to, or large compared with unity. Each of these regimes will be discussed in detail in this section. The solutions also depend on whether the electric vector in the incident wave is perpendicular to (denoted by subscript " J_") or parallel to (denoted by subscript "||") the scattering plane. It is found that the scattering process does not change the direction of polarization. Thus, if the incident wave is polarized perpendicular to the scattering plane, the scattered wave will be perpendicularly polarized also, and similarly for parallel-polarized incidence. However, the scattering coefficients are different for the two directions of polarization. The general behavior of the extinction efficiency as the size of the sphere is varied is illustrated in Figure 5.3 for the case of a real index of refraction, m = n = 1.50. In this figure, QE is plotted versus the parameter (n - 1)X. Because k = 0, Qs = QE, and QA = 0. For small values of X the efficiency increases nonlinearly proportionally to X4 to a value near 2 at (n — 1)X — 1. As (n — \)X increases, QE develops
72
5. Single-particle scattering: perfect spheres
a series of small ripples superimposed on much larger oscillations. The first maximum of the large oscillations is roughly at QE ~ 4 and (n-l)X ~ 2 , followed by a minimum of somewhat less than 2 at (n - 1)X = 3.8. Thereafter, QE oscillates about the value 2 with decreasing amplitude and with successive extrema separated by values of (n — 1)X differing by TT. Figure 5.3. Extinction efficiency of a sphere of refractive index m 1.50 + iO and size parameter X versus (n - \)X.
10
15
20
Figure 5.4. Extinction efficiency of a sphere of refractive index m 1.50 + /0.25 and size parameter X versus (n - l)X.
2
-
10
15
20
5.D. Properties of the Mie solution
73
Figure 5.4 shows the extinction, scattering, and absorption efficiencies of an absorbing sphere with m = 1.50 + 0.25/. This value of k corresponds to an absorption length a" 1 = X/4irk (Chapter 2) equal to the diameter of a sphere with X = l. For small values of X, QE now is proportional to X. As X increases, QE goes through a maximum at (n — \)X — 2 and then slowly approaches QE = 2 with no oscillations or ripples. The fact that the efficiencies can be greater than 1 may be surprising at first sight. Physically, this means that the particle affects a larger portion of the wave front than is obstructed by the geometrical cross-sectional area of the sphere. 5.D.2. X «: / : the Raleigh region When X «c 1 it is useful to expand the Mie expressions for the scattering parameters in powers of X. To terms of order X4,
Im
8
X2m2-lm4+27m2+38
1 + -TT —5— 2 +2 m \ 15 m2 +2
—5—
2m2 +3
m2-! m2+2
(5.11)
and QA = QE - Qs. If k = 0, 8
n2-\ n2+2
(5.12)
QA = 0, and w = 1. In this region the scattered radiance is proportional to I/A 4 , a result that was first obtained by Rayleigh (1871) using dimensional arguments. Thus, particles small compared with the wavelength are known as Rayleigh scatterers. If k is not negligible, but \m\X «: 1, so that higher powers of X can be ignored, then QA~QE =
t 2 l 2 / ^ 2 t u2 (n + k ) +A(n2-k )
r
2
2
2
2
T
2
\(n + k ) +4(n -k ) + 4\
(5.14)
74
5. Single-particle scattering: perfect spheres
In this case w = QS/QA a ^3> s o that if the particle is absorbing, its single-scattering albedo is very small. Small absorbing particles are called Rayleigh absorbers. The absorption efficiency is proportional to X, so that the absorption cross section aAaira2X, and the power absorbed is proportional to the volume or mass of the particle. The intensity of the radiance scattered with the electric vector perpendicular to the scattering plane is proportional to 1, and parallel to the scattering plane is proportional to cos2g. If the incident light is unpolarized, the particle phase function is p(g) = f(l + cos2g). (5.15) Note from equations (5.10) and (5.11) that the denominators of QE and Qs both contain the factor m 2 +2. Thus, the efficiencies can become very large when m2 = Ke ~ — 2, which requires Kei <: 1 and Ker=*—2. Figure 3.1 shows that this can happen in the region of anomalous dispersion on the high-frequency side of a strong absorption band. This phenomenon is known as a plasma resonance, and it can be important in small particles, or asperities that behave as small particles on the surfaces of larger particles, in the infrared. As in Chapter 4, the linear polarization of the radiance is defined as where the subscript " J_" denotes the electric vector perpendicular to the scattering plane, and the subscript "||" denotes the electric vector Figure 5.5a. Phase function of a Rayleigh particle. 1.5
P(9) 0.5
-
30
60
90 g (degrees)
120
150
180
5.D. Properties of the Mie solution
75
parallel to this plane. Thus, the polarization of the scattered radiance is
).
(5.16)
The phase function and polarization are plotted in Figure 5.5. The polarization is always positive, and P = 1 at g = 90°. 5.D.3. X ~ / ; the resonance region
When the particle size is of the order of the wavelength, the behaviors of the efficiencies and phase function are complicated and vary from case to case. If the particle is not too absorbing, the extinction and scattering efficiencies tend to oscillate, with successive maxima or minima spaced such that the quantity (n - 1)X changes by approximately 77 between each. These broad oscillations are an interference phenomenon between the portion of the wave front that passes near the sphere and the portion transmitted through it. The part of the wave propagating outside the sphere undergoes a change in phase equal to lira /A in a distance equal to a, where A is the wavelength in the surrounding medium. The phase of that part of the wave traveling through the sphere changes by nlira/X in this distance. Thus, destructive interference will occur if the difference in phases, (n — i)X, is an odd multiple of ir, and constructive interference if it is an even multiple of v. Figure 5.5b. Polarization of light scattered by a Rayleigh particle.
0.8
0.6 P(g) 0.4
0.2
30
60
90 g (degrees)
120
150
180
76
5. Single-particle scattering: perfect spheres Figure 5.6a. Phase functions of particles of X = 1 with m = 1.50 + /0 (solid line) and m = 1.50 + /0.25 (dashed line). The single-scattering albedo of the particle with k = 0 is w = 1.00, and that of the particle with k = 0.25 is w = 0.27. 2.5
x == 1 n = 1.50
2
-
1.5
-
1
-
0.5
-
-
/
P(9) /
~
\ \k =-
]
0.25 I
r\
30
60
90 120 g (degrees)
l
150
l
"
180
Figure 5.6b. Polarization of light scattered by particles of X = 1 with m = 1.50 + IO and m = 1.50 + /0.25.
0.8
0.6
0.4
0.2
30
60
90 g (degrees)
120
150
180
5.D. Properties of the Mie solution
11
The smaller ripple structure superimposed on the broad oscillations is also obviously some sort of interference phenomenon, although its origin is obscure. It is thought to be caused by radiation in modes that travel close to the surface of the particle. The particle phase functions and linear polarizations for the cases m = 1.5 and m = 1.504-/0.25, with X = l, are shown in Figure 5.6. Note that the phase function has become forward-scattering, but that the polarization is similar to that of a small particle. 5.D.4. X » 1: the geometric-optics region 5.D.4.a. Introduction The phase functions and polarizations of a nonabsorbing sphere (m = 1.50+/0) and absorbing sphere (m = 1.50+/0.25) of Z = 100 are shown in Figure 5.7. When the sphere is large compared with the wavelength, the parts of the wave front separated by distance of the order of the radius are nearly independent, so that what happens to one part has very little effect on the other parts. That is, to a large extent the propagating wave can be treated as if it were made up of independent bundles of rays that are refracted and reflected by the sphere. With certain exceptions, the features of Figure 5.7 can be understood in terms of Figure 5.7a. Phase functions of particles of X = 100 with m = 1.50 + iO (solid line) and m = 1.50 + /0.25 (dashed line). The single-scattering albedo of the particle with k = 0 is w = 1.00, and that of the particle with k = 0.25 is w = 0.54. 100
10
p(g)
0.1
r
0.01 30
60
90 120 g (degrees)
150
180
78
5. Single-particle scattering: perfect spheres
geometric optics. The various phenomena that control different portions of the curves of this figure will now be discussed in detail. 5.D.4.b. Diffraction At large values of X the extinction efficiency approaches the asymptotic value of 2, not 1, indicating that the sphere affects a portion of the incident wave front equal to twice its geometric cross section. Also, the phase curves of both the clear and absorbing particles exhibit a very strong, narrow peak in the forward-scattering direction. Both of these effects are due to diffraction and follow from the wave nature of light. In order to understand this phenomenon, we shall first take a brief digression and consider the scattering of light by a circular hole in an opaque wall. The situation is shown in Figure 5.8. A plane wave of irradiance / is incident on a hole of radius a in an opaque wall of infinite extent oriented perpendicular to the direction of propagation. A white screen is located a distance L from the hole, where L is large compared with both a and A and also with a2/A. The question of interest is the pattern of the electric fields and the resultant intensity at the screen. According to Maxwell's equations, a changing electric field generates a magnetic field, and a changing magnetic field generates an electric field. Thus, a propagating electromagnetic wave may be regarded as generating itself; that is, the changing fields on a wave front Figure 5.7b. Polarization of particles of X = 100 with m = 1.50 + iO (solid line) and m - 1.50 + i0.25 (dashed line).
0.5
P
•
0
-0.5
"
-1
30
60
90 120 g (degrees)
150
180
5.D. Properties of the Mie solution
79
generate succeeding parts of the wave. This concept leads to Huygens's principle, in which each point on a wave front may be considered to be a source of wavelets that travel radially outward and combine coherently with wavelets from other points to produce new wave fronts. If the wave front is plane and infinite in lateral extent, this process simply produces another plane wave front. If part of the wave front is obstructed, Huygens's principle may be used to calculate the resultant fields elsewhere. Let the electric field at the hole be Ee0, which is related to the incident irradiance / by equation (2.24), J = y/ee0 /fim0 Ej0 / 2 . The intensity of a Huygens wavelet generated by the incident field at a given point in the hole falls off inversely with the square of the distance from that point, so that the field strength of the wavelet falls off as the reciprocal of the first power of the distance. Hence, the fields at a point &>f located at position (jt',y',z') = (£cos
•dA',
where • = [ ( * - x') 2 + (y - y') 2 +(z - z') 2 ] 1/2 . Figure 5.8. Schematic diagram of the diffraction by a hole.
80
5. Single-particle scattering: perfect spheres
Because a
2 ' + / + zz') C
2*
Because I is not very different from ^, ^ may be replaced by r in the denominator of the integral. However, because I » A, small changes in I could make a large difference in the phase, so that the first-order terms must be retained in the exponent. Because 0 2/A#"
These integrals may be evaluated using the well-known relations (Jahnke and Emde, 1945)
J}(u) = f e'<« cos
(5.17)
and
where /^(M) is the Bessel function of the first kind of argument u and order / . Putting / = 0 and u = 2ir(£/A)sin0, and using the fact that / 0 ( - u) = J0(u), gives Ee = Ee0^
= F
r2Tri(r/\-vt)
u
where X=2ira/X. Hence, the diffracted power per unit area falling on the screen at &> is
f (5.19)
5.D. Properties of the Mie solution
81
The constant ^ is found by requiring that the total power falling on the entire screen be equal to the power coming through the hole, lira2 = Tra2^/ee0 / fjLm0E20 /2; or, because sin0 = Jt//s and dA = 2rrxdx,
where v = Xx/r. This integral can be evaluated using (5.18) along with another recursion relation involving the Bessel functions (Jahnke and Emde, 1945):
,
]
%
)
.
(5.20)
Putting / = 0 in (5.20) gives
J1(v)/v =
J0(v)-dJ1(v)/dv,
so that, from (5.20), with/ = - 1 ,
Because / 0 (0) = 1 and Jx(0) = /0(oo) = jt(oo) = 0, the integral has the value j , and the total power falling on the screen is Jira2 =
Jva2(^a/X)2,
so that g7 = X/a. Hence,
XsinO
'
and the power per unit area in the diffracted light pattern at distance /* from the center of the hole is
dA~~Ja
J a
2
*
,22
where a = ira 2 is the area of the hole. Because the solid angle dil subtended at the center of the hole by dA is dCL = dA/r2, the
82
5. Single-particle scattering: perfect spheres
diffracted power per unit solid angle is dP£=dPldA_ T o J da dA d£l~ 4ir
\ \l
(5.23)
Xsinfl
The function 2JX(Z)/Z, where Z = Xsin0, and its square are graphed in Figure 5.9. The function has the value of unity at Z = 9 = 0 and goes through zero at Z = 1.22TT; thereafter it oscillates with decreasing amplitude about zero at intervals of Z separated apFigure 5.9a. Diffraction function 2/ 1 (z)/z and its square.
0.8
"
0.6
0.4
-
0.2
"
0
"
-0.2
10
15
Figure 5.9b. Enlargement of part of Figure 5.9a. 0.1
—
0.08
"
0.06
"
0.04
"
0.02
"
10
15
5.D. Properties of the Mie solution proximately by TT. Thus, the intensity pattern of the light transmitted through the hole does not have a sharp edge, as would be the case if geometric optics held, but consists of a bright central peak surrounded by a series of alternate dark and bright rings. The angular width of the central peak is inversely proportional to the radius of the hole. The electric field and the intensity are indicated schematically in Figure 5.10a. The field consists of two parts: the incident field and the diffracted field. These can be separated experimentally by using a detector with a directional response instead of the white screen that glows independently of the direction of arrival of the radiation. As indicated in the figure, the incident radiation arrives from the vertical direction, and a directional detector that looks exactly upward will respond only to the incident irradiance. However, the diffracted intensity appears to come from the direction of the hole, so that when the detector is pointed toward the hole it will measure only the scattered radiance. Now consider the complementary situation. Remove the wall and replace it by an isolated opaque disk of the same size and at the same position as the hole. The electric field and intensity on the screen are indicated schematically in Figure 5.10b. The field in the diffraction pattern of the disk can be found from Babinefs principle: the hole and the disk are exactly complementary. If the hole is plugged by the disk, all fields must disappear. Hence, the field diffracted by the disk is exactly equal to the negative of the field diffracted by the hole. Because the intensity is proportional to the square of the field, the power per unit area in the diffraction pattern of the disk is exactly the same as that of the hole and is given by (5.21). The total light on the screen is the sum of the incident irradiance / and the diffracted power dPd/dCt. This is shown in Figure 5.10b. The two components may, in principle, be separated, as illustrated in Figure 5.10c. The diffracted component of the scattered light does not depend on the nature of the disk, or even the details of its shape, but only on constructive and destructive interference of the incident wave front in the vicinity of the disk. Thus, the power diffracted into unit solid angle by a sphere of radius a is the same as that diffracted by a circular hole of the same radius and is given by (5.23). Comparing (5.23) with (5.4), it is seen that as far as the diffracted light is concerned, a sphere behaves like a particle that scatters light with efficiency Q s = 1 and
83
84
5. Single-particle scattering: perfect spheres Figure 5.10a. Schematic diagram of the electric field and intensity in the diffraction pattern produced by a hole in a wall of infinite extent. Incident light
Diffracted field
Diffracted radiance
Figure 5.10b. Schematic diagram of the electric field and intensity produced by a disk. The total field consists of the sum of the incident and diffracted fields. Incident light
Incident electric field or irradiance
Combined incident and diffracted fields
Combined incident and diffracted radiance
5.D. Properties of the Mie solution
85
diffractive phase function
pd(g) = X2[2J1(Xsin0)/XsinO]2 = X {2/ 1 [^sin(7r — g)] / X sin(ir — g)} .
(5.24)
The height of the diffraction peak is X2, and its angular half-width at half-maximum A0 is given to a good approximation by Zsin(A0)~ TT/2, or A large particle may be considered as affecting an area of the wave front approximately equal to twice its geometric cross section. The portion of the wave that actually encounters the particle is absorbed or scattered by refraction and reflection, while the portion passing between the particle surface and a distance about yfla from the center contributes most of the diffracted light. This is the reason that a large particle has an extinction efficiency QE « 2, instead of 1, and a phase function with a high, narrow peak in the forward direction. Figure 5.10c. Schematic diagram of the separated incident and diffracted fields from a disk. Incident light
Incident electric field or irradiance
Diffracted radiance
\
I Diffracted field
86
5. Single-particle scattering: perfect spheres
For a nonconducting particle the scattered intensity is the same for both directions of polarization (Liou and Hansen, 1971). Hence, if the incident irradiance is unpolarized, the diffracted radiance is also unpolarized. 5.E. Geometric-optics scattering 5.EJ. Scattering not caused by diffraction
A little thought will show that the discussion of the preceding section is valid only for a large, isolated particle. The diffraction can be associated with scattering by the particle because constructive interference of Huygens wavelets from light that passes through a complete annulus around the particle generates a wave that appears to radiate from the particle into a narrow cone of approximate angular width 77 /X about the forward direction. However, this book is concerned with soils and powders in which there are large numbers of particles so close together they touch. Consider a screen made of a monolayer of nonabsorbing spheres just touching one another. In this case, parts of the annulus around each particle are blocked, altering the nature of the interference. Because only the light that passes through the spaces between the particles can contribute to the diffraction, the diffracted light must be regarded as being associated with the holes and not with the particles. Furthermore, if the medium is a powder, the light diffracted by the holes between the particles travels very close to the forward direction and encounters more particles within a very short distance. Thus, the angular deviation caused by the diffraction is undetectable in practice, and collimated light that has passed through the holes between the particles is still collimated for all practical purposes. This point has not been widely appreciated. An observation supporting the assertion that diffraction is not applicable to closely spaced particles is the brightness of the new moon. Lunar soil contains a large number of particles of the order of 10 fim in size. When isolated, these would have a diffraction pattern of half-width of the order of 3°. If such diffraction were important on the lunar surface, the illuminated portion of the moon should be much brighter than the full moon when less than about 3° from new. However, the moon is invisible when less than about 7° from new (Hapke, 1971).
5.E. Geometric-optics scattering
87
Denote the portion of the scattering efficiency that is associated with diffraction by Qd. If the particle is isolated, Qd = 1. However, when the particles are so close togethet that they nearly touch or actually touch, as in a powder or soil, effectively Qd = 0. Hence, it is useful to explicitly separate the part of the scattering that is associated with diffraction from that due to other effects. The nondiffractive portion of the scattering efficiency will be denoted by Q with a lowercase subscript s and defined by Qs = Qs~Qd(5-25) Except in certain directions, the nondiffrdctive interactions of the sphere with the incident wave included in Qs can be described reasonably well by geometric optics. These interactions consist of rays that are specularly reflected from the surface and rays that are transmitted through the sphere. The latter may be internally reflected one or more times from the inside of the surface. The major phenomena are shown schematically in Figure 5.11. It is important to note that because of the symmetry of the sphere, the rays that are reflected or refracted by the sphere always remain in the same plane as that formed by the incident ray of interest and the axis of the sphere parallel to the incident rays. 5.E.2. Surface reflection
If the sphere is large, the portion of the surface encountered by a ray can be considered as locally flat and effectively infinite, and the Fresnel equations may be used to calculate the amounts and directions of the light reflected and transmitted. The geometry of the surface-reflected light is shown in Figure 5.12. Figure 5.11. Schematic diagram of the scattering processes in a sphere with X » 1. surface-reflected ray
twice-transmitted, once y internally-reflected ray
^ ^ ^
j
^sconce-transmitted ray
diffracted ray
88
5. Single-particle scattering: perfect spheres
A sphere of radius a is illuminated by irradiance / ± u, where the subscript " ± II" is to be interpreted as " _L " if the incident irradiance is polarized with the electric vector perpendicular to the scattering plane, and as "||" if the electric vector is parallel to the plane. The rays are incident at points where the radius vector makes an angle # with the direction of propagation. For clarity only one such ray is shown; however, the rays are distributed symmetrically in azimuth all around the axis of the sphere. These rays are reflected through a phase angle g = 2#. Consider a second set of rays incident at points making angles # + d # with the radius vector and reflected through a phase angle g + dg = 2 # +2d&. The area of surface between all those rays whose intersections with the sphere are bounded by the angles # and ft + dft is dA = 2Trasind'adft. Because the projection of dA onto the plane perpendicular to the direction of incident irradiance is dA cos #, the power dP6 specularly reflected by dA is
cos
2
Jl\R (5.26)
where / ? ±
and y/
are the Fresnel reflectivities defined in
Chapter 4. At a large distance from the sphere this light passes through an incremental solid angle dft Figure 5.12. Specular reflection from the surface of a large sphere.
5.E. Geometric-optics scattering
89
Hence, the power per unit solid angle specularly reflected from the surface of the sphere through an angle g is y J
R±i(g/2)
The curvature of the spherical surface has a defocusing effect, which causes the radiation incident on the sphere through a range of angles d& to be reflected through a range of angles Idd. Because R ± > Rp if the incident irradiance is unpolarized, the light reflected from the surface always has a net positive polarization. The total amount of light reflected from the surface of the sphere is, from (5.26), ffi rir/2
'0
If the incident light is unpolarized, so that J± = J^ = / / 2 , then the contribution to Qs and p(g) from surface reflection is e,P,(g) = [i? ± (g/2)+i?n(g/2)]/2, (5.29) where Qs = S e ,
(5.30a)
and Se is the total fraction specularly reflected of the light that is externally incident on the surface of the particle, (5.30b) 5.E.3. Refracted rays
The remainder of the light scattered by the sphere consists of rays that are refracted into the sphere and back out after making one or more internal traverses of the particle. These rays are shown in Figure 5.13. In this section it may be assumed that k <: 1, because otherwise the particle would be opaque and the contribution of the refracted rays would be negligible. Consider first a ray of irradiance /±n that makes only one pass through the sphere after encountering the surface at the point where the radius vector makes an angle # with the direction of propagation. A fraction R±\\(ft) is reflected by the first surface, and the remainder
90
5. Single-particle scattering: perfect spheres
enters the particle as [1 — R±\\(/&)]cosft/cos/&\ making an angle &' with the radius vector, where #' is given by Snell's law [equation (4.24)]. As the ray traverses the sphere it is attenuated by absorption according to exp( - ax) = exp( - AkXcos #'), where a = 4 irk /A is the absorption coefficient and x = 2 a cos #' is the path length through the sphere. Because k <^ 1, equations (4.22) and (4.23) may be used to calculate R±0). At the far side of the sphere a fraction R±\\(&') of the ray is reflected. It may readily be shown from the symmetry of equations (4.22) and (4.23) that the reflection coefficient for rays incident internally on a surface at angle ft' is the same as the reflection coefficient for rays incident externally at angle ft. Hence, at the forward surface, a fraction RL\\{ft) of the energy is reflected, and fraction [1 — ^il|(#)]cost?Vcos# is transmitted. At all positions the ray remains in the same plane. Figure 5.13 shows that the once-transmitted ray exits the sphere in a direction making an angle T = 7 r + 2 # - 2 # ' with the direction to the source. The phase angle of the exiting ray is g = 2ir - T = TT — #'). The minimum value of g occurs when # = TT/2, SO that
Figure 5.13. Internally refracted and reflected rays in a large sphere. Y
5.E. Geometric-optics scattering
91
#' = #c = sin'Hl/n), or g = 77 - 2 ( T T / 2 - # C ) = 2#c . Hence, the angular width of the forward-refracted peak is ir~2# c , where ftc is the critical angle for total internal reflection. Of the power incident on the sphere at all points on the surface between angles ft andft+ dft through the area dA cosft=* lira sinfta dft cosft,an amount emerges between angles g and g + dg, where the subscript 1 on dP indicates the power emerging after one transit of the sphere. At a large distance from the particle this light passes through an incremental solid angle dil = 2ir sin gdg. Hence, the power per unit solid angle of light transmitted once through the sphere is dil
~~
R±ll(ft)]2e-4kXco*»'smftcosftdft 2TT sin
'
gdg
The remaining power is internally reflected, passes through the sphere a second time, and is partially transmitted out of the sphere; the rest is again internally reflected, and the process continues. Because of the particle symmetry, at each reflection ft' and x are the same, and the ray stays in the same plane. From Figure 5.13 it is seen that light emerging from the sphere after each subsequent internal reflection and transmission differs from dP1±\\/d£l in three ways: (1) each internal reflection attenuates the intensity by R±\\(&); (2) each transmission attenuates the intensity by e-4kXcos^f-<) (3) each internal reflection increases the deflection angle T by an additional amount 77 - 2 # ' . hence, rays that have passed through the particle ^ times and scattered into phase angle g have power per unit solid angle
(5.32a) where Ff(g, &,•&') is the focusing factor, 4 sin & cos # sin g(dg/d&)
4 sin # cos # , smT(dT/d&)
(5.32b)
-fir+2d-2f0\
(5.32c)
= \T-2J^v\,
(5.32d)
92
5. Single-particle scattering: perfect spheres
where JV is the integer (J^ > 0) that makes 0 < g < ir; # and #' are related by Snell's law, sin # = n sin #', so that dT/d& = 2 ( 1 - fdV/dti)
= 2 { l - ^[(l-sin 2 #)/(rt 2 -sin 2 #)] 1 / 2 }. (5.32e)
Equations (5.32) predict that the intensity becomes infinite at certain angles where rays incident on the sphere over a range of entrance angles # are focused into the same direction T. This is another situation where geometric optics breaks down. Although the particle phase function is sharply peaked at these angles, the actual intensity remains finite, and its value must be calculated by Mie theory. However, ray theory does correctly predict the angles at which the peaks occur. According to equation (5.32e), a peak occurs when dT/d& = 0, or
sin# = [ ( 9 2 - « 2 ) / ( ? 2 - l ) ] V 2 .
(5.33)
The ray corresponding to ^ = 2 (i.e., two transits plus one internal reflection) is responsible for the primary rainbow in water droplets. Taking the refractive index of water in visible light a s n = 1.33, these equations show that the rainbow ray enters the sphere at # = 59.4°, with #' = 40.2°, and exits at deflection angle T = 318° or phase angle g = 42°. Fainter rainbows that are produced by rays making more than two transits of the sphere may also be observed. For substances other than water the generic term cloudbow is sometimes used instead of rainbow. Another situation in which a singularity in FJig ,&,&') occurs is if g = 0, but # # 0 or 77/2. The resulting peak is called the glory. (A glory can frequently be seen from an airplane flying above a cloud of water droplets as a bright glow around the shadow of the aircraft on the cloud. However, in water the refractive index is too small for the ^ = 2 ray to cause a glory, which in this case is thought to arise from rays with ^ » 2 that enter the sphere at nearly glancing incidence and are multiply internally reflected or guided around the surface of the particle.) Because of the symmetry of the particle, the rays always remain in the same plane during the processes of reflection and refraction by the various surfaces. Hence, the total contribution of the refracted
5.E. Geometric-optics scattering
93
rays to the scattering is the sum of all terms of the form of (5.32), _ ^
dQ, ~ L
9
dil
(5.34) In order to illustrate the behavior of the refracted rays, the ^ = 1 and ^ = 2 terms will be discussed in detail. When ^ = 1, T = ir + 2 ( # - # ' ) . For the axial ray, # = 0, T = g = 7r, and ^(77,0,0) = (1 — 1/n)" 2 . As # increases, T increases to a maximum value of T = 2TT — 2 ^ , where t^ = s i n ' K l / w ) is the critical angle for total internal reflection, corresponding to the ray that enters the sphere at # = TT/2. At the same time, g decreases from IT to 2#c'. The result is a bright lobe of forward-refracted light. The sphere acts as a thick lens to focus the incident light into a beam of angular half-width 2 ^ . When 9 = 2, T = 2TT + 2 # - 4 # ' . For the axial ray, # = 0, T = 2TT, g = 0, and 772(0,0,0) = (l — 2/n)~2. As # increases, T first decreases, and g increases to the cloudbow angle. As # continues to increase, the rays are prevented by Snell's law and by the curvature of the sphere from exiting at any lesser deflection angle T, but instead pile up, causing the cloudbow. T then increases to a maximum angle of 3TT —4sin" 1 (l/n) corresponding to # = ir/2 and sin#' = 1/n. If n is large enough, this maximum T may exceed 2TT. In this case there will be a glory when T = 2ir; according to (5.31c), this occurs when # = 2#', or sin # = n(l - n2/4)1/2.
5.E.4. Total light scattered by a large sphere Combining the equations developed in this section, the total power, including diffraction, scattered into unit solid angle by an isolated sphere, large compared with the wavelength, is
dp±l] /da = dPd/dn + dpi±ll /da + dP,±,,/
94
5. Single-particle scattering: perfect spheres
(5.23), (5.27), and (5.34). Hence,
2J1[Xsm(7r-g)}\2
-g)
2
I
X L R^W'e-'^^'F^g^,^').
(5.35)
The scattering efficiency and phase function may be separated by integrating [Qsp(g)]J.ii over all solid angles:
fj
(5.36a)
the particle phase function is (536b) As a check on computations, Qs must equal 2 when k = 0, because the only loss occurs by absorption of the refracted rays. Approximately 99% of the light scattered by a large nonabsorbing particle is included in the terms up to and including the second transit of the particle. Hence, the infinite sums in (5.34) or (5.35) may usually be truncated at ^ = 2. If the particle is absorbing, it may be necessary to include only the first transmitted term. If the incident irradiance is unpolarized, so that J± = /y = / / 2 , then {Q
^,-&,-&').
(5.37)
If Qs is divided into (1) a strongly forward-scattering term associated with diffraction and (2) the rest, which scatters light over a broad range of phase angles, Qs = Qd + Qs, then Qd = 1 or 0, depending on
5.E. Geometric-optics scattering
95
whether or not the particle is isolated, and
(5.38)
9
The phase function of an isolated sphere with m = 1.50 + 0/ and A" = 100 calculated from geometric optics is shown in Figure 5.14a, where the different contributions are shown explicitly. Comparing this figure with the exact Mie-theory calculation shown in Figure 5.7, it is seen that the two results are qualitatively similar. Although significant differences exist, closer inspection of the figures shows that most of these differences are associated with narrow resonance peaks and that the overall agreement is good. Hodkinson and Greenleaves (1963) and Ungut, Grehan, and Gouesbet (1981) have shown that in the forward-scattering portion of the phase diagram where the scattering angle is smaller than 20° there is good agreement between Mie theory and geometric optics for X as Figure 5.14a. Phase function of a large sphere with m = 1.50 4- /0 calculated according to geometric optics. Compare the total intensity with the solid curve of Figure 5.7a. The two curves are similar except for the resonances in the Mie solution. 100
I
• I ' I ' I
^
^
I '
I ' I
m = 1.50 + iO 10
Tota p(g) internally /reflected 0.1
0.01
30
60
90 120 g (degrees)
150
180
96
5. Single-particle scattering: perfect spheres
small as 6. However, according to Liou and Hansen (1971), satisfactory agreement over the complete range of phase angles is not reached until X exceeds about 400. We are now in a position to understand the various parts of Figures 5.7a and 5.14a. The sharp, narrow peak at g = IT is due to diffraction. The broad forward-scattered lobe is primarily due to the refracted ray that has made one traverse of the sphere. The backscattered lobe, the rainbow at 21°, and the glory are all due to the internally scattered rays. Note that there is a gap between the maximum value of g at which the internally reflected ray emerges and the minimum phase angle at which the forward-refracted ray emerges. In this gap, the only contribution to the scattered light is from surface Fresnel reflection, so that p(g) is small there. The diffracted light is independent of k. If k «: 1, the surface reflectivity is nearly independent of k. However, the intensities of the refracted rays are exponentially dependent on k. As k increases, the refracted component decreases, and the internally reflected lobe scattered in the backward direction decreases more rapidly than the once-transmitted forward-scattered light. If kX » 1 , but k
The linear polarization for unpolarized incident irradiance may be calculated from
where [QsP(§)]±\\ a r e given by (5.35). The polarization for a sphere with m = 1.50+i0 and Z = 100 is shown in Figure 5.14b. The light reflected from the surface is positively polarized, the once-transmitted light is negatively polarized, and the internally reflected light is positively polarized. The cloudbow is positively polarized, while the glory may have positive or negative polarization. Thus, the polarization is positive for small and intermediate phase angles and negative for large phase angles.
5.E. Geometric-optics scattering
97
5.E.6. The equivalent-slab approximation for Qs
When the particle is large compared with the wavelength and the incident light is unpolarized, an analytic approximation for the nondiffractive component of the scattering efficiency that is sufficiently accurate for most applications may be derived by replacing the sphere by a slab with appropriate optical properties. The equivalent-slab model is illustrated schematically in Figure 5.15. Figure 5.14b. Polarization of a large sphere with m = 1.50 + I'O calculated according to geometric optics.
0
20
40
60
80
100 120 140 160 180
Figure 5.15. Schematic diagram of the equivalent-slab model for Qs.
(1-5 e )0
(1-S e )0 4 5 j 3 (1-S i )
(i-s e )e 3 s 1 2 a-s 1 )
98
5. Single-particle scattering: perfect spheres
We have seen that the total fraction of incident light specularly reflected into all directions from the outer surface of a sphere is Se, where Se is the integral of the Fresnel reflection coefficients, equation (5.29). Thus, for externally incident light, the reflection coefficient of the surface of the equivalent slab is set equal to Se. A fraction 1 — Se of the light enters the slab and is then attenuated by absorption. Define the internal-transmission factor 0 as the total fraction of light entering the particle that reaches another surface after one transit. From (5.31) and the discussion preceding this equation, for a sphere this factor is given by
0 = 2 r/2e-4kXcos»'sin
#cos
From Snell's law, sin # cos &d& = n2sin tf'cos &'d&'. Also, AkX = aD, where a is the absorption coefficient, and D = 2a is the diameter of the sphere. Hence, 'cos #'#', where ^ = sin'Kl/n).
This integral is readily evaluated to give
(5.40) Define the mean ray path length (D) as the thickness of a slab that will have the same value of 0 as a -> 0. For the slab, 0 = e~a
0 - 1 - f [n2 - (l/n)(n2 - l)^2]aD. Hence,
V2
]z).
(5.41)
This quantity is the average distance traveled by all rays during a single transit of the particle. The value of (D)/D ranges from about 0.85 when n = 1.3 to 0.93 when n = 2. By definition, when a
(5.42)
99
5.E. Geometric-optics scattering
where (D) is given by (5.41), and the thickness of the equivalent slab is set equal to (D). Following the first passage through the particle, a fraction 5,- is internally reflected, and the remainder 1 - St is refracted through the surface, where St is the reflection factor for light internally incident on the surface of the particle. The process continues, as indicated schematically in Figure 5.15. Because of the symmetry of the sphere, the reflection coefficients are the same for each order of internal reflection. Thus, the total fraction of incident light that emerges from the particle is Q, = Se + (1 - 5J@(l - 5,) + (1 - SJBSflil
- S,)
+ ( i - s e ) e s I . e s / e ( i - s I . ) + ••• (5.43)
Because of the spherical shape of the particle, St = Se. Hence, for a sphere, (5.44) Figure 5.16. Scattering efficiency of a large sphere versus aD calculated according to geometric optics and from the equivalent-slab model, equation (5.44); m = 1.50 4- iO. The two curves are indistinguishable on the scale of thisfigure.The maximum error is .002. n = 1.50 0.8
0.6
0.4
0.2
4
6
aD
10
100
5. Single-particle scattering: perfect spheres
where ® is given by (5.42). Equation (5.44) is the equivalent-slab approximation for Qs. The scattering efficiency versus a(D) calculated using the exact geometric-optics expressions (5.38) and the approximate expression (5.44) is shown in Figure 5.16, where the approximation is seen to be excellent. To a similar accuracy the absorption efficiency is
QA-l-Q9~(l-St)£^g.
(5.45)
Note that when a( D) <^ 1, QA - a( D) oc aD, so that the total amount of light absorbed J
The general principles on which Mie theory is based can be extended to other relatively simple geometries, including nonuniform, spherical particles in which the surfaces of discontinuity in the refractive index are concentric spheres, right circular cylinders of infinite length, and ellipsoids of revolution. For cylinders, the size parameter refers to the radius of the cylinder. The equations for the coated spheres and cylinders are derived in the standard references, including those by Van de Hulst (1957), Kerker (1969), and Bohren and Huffman (1983). The Bohren and Huffman book contains FORTRAN programs for calculating scattering by a coated sphere and infinite circular cylinder. Scattering by oblate and prolate ellipsoids of revolution is treated in Asano and Yamamoto (1975). The properties of the solutions are qualitatively similar to those of a sphere with similar size parameter X. Although different in detail, the scattering exhibits Raleigh behavior when X is small, and diffraction, interference, and cloudbow-like resonance phenomena when X > 1.
6 Single-particle scattering: irregular particles
6.A. Introduction
The scattering of electromagnetic radiation by perfect, uniform, spherical particles was described in Chapter 5. However, such particles are rarely found in nature. Most pulverized materials, including planetary regoliths, volcanic ash, laboratory samples, and industrial substances, have particles that almost invariably are irregular in shape, have rough surfaces, and are not uniform in either structure or composition. Even the liquid droplets in clouds are not perfectly spherical, and they contain inclusions of submicroscopic particles around which the liquid has condensed, so that they are not perfectly uniform. At the present state of our computational and analytical capabilities it is not possible to find exact solutions of scattering by such particles, so that it is necessary to rely on approximate models. The objective of any model of single-particle scattering is to relate the microscopic properties of the particle (its geometry and complex refractive index) to the macroscopic properties (the scattering and extinction efficiencies and the phase function) that, in principle, can be measured by an appropriate scattering experiment. This chapter describes a variety of models that have been proposed to account for the scattering of light by irregular particles. This is not an exhaustive survey; rather, it is a commentary on those models that are most often encountered in remote-sensing applications or that offer some particu101
102 6. Single-particle scattering: irregular particles lar insight into the problem. Finally, a model based on recent experimental work is presented. 6.B. Extension of definitions to nonspherical particles
The physical meanings of the particle scattering parameters defined in Chapter 5, including the cross sections and phase functions, are intuitively clear for a particle that is spherically symmetric. However, if a particle is not spherical, the proper interpretation of these quantities is not obvious. Unless explicitly stated otherwise, in this book it will always be assumed that we are dealing with ensembles of particles whose orientations are random. Thus, the power extinguished by a particle of arbitrary shape is defined to be the average of the power removed from the beam of light as the particle is randomly oriented in all directions. Similar definitions apply for the power scattered and absorbed. The geometric cross section a of a particle is the average area of the geometric shadow cast by the particle as it is oriented at random in all directions. The equivalent particle radius and diameter are defined as (6.1a) (6.1b) With this understanding of implicitly averaging over all orientations, the quantities defined in Section 5.B have physical meanings for nonspherical particles, as well as for spherical ones. These meanings are intuitively clear and internally consistent. Note that the scattering properties defined in this way are azimuthally symmetric, so that p(g) is independent of azimuth angle, and (5.6b) holds. In some references the effective particle radius a is defined to be the radius of a sphere of the same volume as the particle under study. With such a definition the power absorbed by the particle is the same as that absorbed by the equivalent sphere if the particle is optically thin, but not if the particle is opaque. In this book the definition of the equivalent radius a in terms of a sphere of the same mean geometric cross-sectional area is preferred, because this leads to unambiguous meanings for the efficiencies and phase functions. Also, the power absorbed by an optically thick irregular particle and that absorbed by its equivalent sphere are equal.
6.C. Extrapolations from Mie theory
103
6.C. Extrapolations from Mie theory 6.C1. Particles small compared with the wavelength
Although Mie theory is strictly valid only for spherical particles, it still serves as a valuable starting point in discussing scattering by irregular particles. Suppose that the condition \m\X
(6.2)
If \m\X
104
6. Single-particle scattering: irregular particles
will scatter light like a sphere, resulting in equations similar to (5.10) and (5.11) for a sphere. That is, QE ~ QA a kX if k is not zero, and Qs oc X4. Hence, electrostatic theory may be employed to calculate pe0, and the result may be used in (6.3) to calculate Qs. The advantage of this method is that it allows a wider variety of particle shapes to be treated analytically than does electrodynamic theory. In particular, scattering by oblate and prolate ellipsoids of rotation may be calculated, which includes disks and needles as special cases, and also triaxial ellipsoids. Scattering by small ellipsoidal particles is discussed in considerable detail by Bohren and Huffman (1983). The net result is that if a particle is equant, that is, if its dimensions are roughly the same in all directions, then its efficiencies are not very different from those of a sphere of similar size. However, if the particle is not approximately equidimensional, the particle still scatters and absorbs in a dipole-like manner, but its efficiencies differ quantitatively from (6.4). 6.C.2. Particles comparable to the wavelength
A number of experimental studies have been carried out on irregular particles with X ~ 1 in visible light, including those by Hodkinson (1963), Holland and Gagne (1970), Pinnick, Carroll, and Hofmann (1976), and Perry, Hunt, and Huffman (1978), and on model particles using microwaves, including studies by Greenberg (1974), Zerull and Giese (1974), Zerull (1976), and Schuerman et al. (1981). Greenberg (1974) also gave equations for calculating scattering by cylinders of finite length and by spheroids. Those studies showed that for equant particles the observed scattering is similar to that predicted by Mie theory for a perfect sphere. Figure 6.1 compares the scattering from a cube with X = 3.75 with the predictions of Mie theory for a sphere of the same size. In the preceding section it was argued that if the phase change across the particle is small, the shape has a relatively minor influence on the scattering. Evidently the same principle applies to departures from sphericity. If an equant particle is not much larger than the wavelength, then the difference in any direction between the particle and a sphere of the same average geometric cross section is small compared with the wavelength; the phase changes across those differences and the resulting effects on the scattering are negligible.
105
6. G Extrapolations from Mie theory
These conclusions were supported by experiments (Zerull, 1976) showing that Mie theory accurately predicts the scattering properties of spherical particles with X » 1 but with surfaces whose scales of roughness are small compared with the wavelength. On the basis of those studies Pollack and Cuzzi (1980) suggested that Mie theory may be used to calculate the scattering and absorbing properties of equant, irregular particles with X < 5. However, no adequate computational procedure exists if the departures from equidimensional shape are large in this size range. Several approaches to this problem have been discussed by Schuerman (1980). A major difference between the scattering properties of irregular particles and perfect spheres is the absence of the large peaks and valleys observed in the extinction and scattering efficiencies of spheres as the size parameter is changed. This can be understood because the varying dimensions in different directions of the irregularly shaped particles and the finite distribution of particle sizes in most real media average out the interference effects that cause these resonances. A study by Hodkinson (1963) of the extinction of visible light by colloidal Figure 6.1. Measured angular scattering coefficient of a cube of X = 3.75 and m = 1.57 + /0.006 (circles), compared with the predictions of Mie theory for a sphere of the same size (lines), for the two polarized components of incident radiation. From Zerull (1976), courtesy of Friedrich Vieweg & Sohn. 1000'
1000
Perpendicular
100
100 —
10-
Parallel
-
10-
1 -
0.1 -
0.1-
.01 -
.01 -
I
90
180
0
Scattering Angle (degrees)
I 90
180
106
6. Single-particle scattering: irregular particles
suspensions (Figure 6.2) showed that, in contrast to perfect spheres, the extinction efficiencies of irregular particles increase smoothly and monotonically with X and level off at QE ~ 2 for (n - 1)X > 3. 6.C.3. Particles large compared with the wavelength
When irregular particles are much larger than the wavelength, Mie theory provides only a rough guide to the scattering properties of irregular particles, and there are major differences between the behaviors of spherical and irregular particles. One reason for this difference can be seen from equation (5.41), which shows that
Measurements on quartz particles
9.6
12.8
6.D. Empirical scattering functions
107
be discussed later in this chapter. Before discussing them in detail, we will describe several empirical scattering functions that are frequently encountered in the literature. None of these has a particularly profound theoretical basis; rather, they are simply convenient mathematical expressions. We will also summarize some of the theoretical approaches that have been used to attempt to model the behavior of large particles. 6.D. Empirical scattering functions 6.D.I. Legendre polynomial representation of p(g)
It is often convenient to represent p(g) as a series of Legendre polynomials: j
(6.5)
7= 0
where the b/s are constants, and the Pj(g)9s are Legendre polynomials of order j . The properties of these functions are reviewed and illustrated in Appendix C. This representation of p(g) is most useful when the departures from isotropic scattering are not very large, so that only a few terms are necessary. It has been used by many authors, including Hapke and Wells (1981) and Mustard and Pieters (1989). From the normalization condition on p(g), equation (5.6), the series must satisfy
But, from the orthogonality of the Legendre polynomials, the only nonvanishing integral is that for j = 0, which equals 4irb0. Hence, bQ = 1. No adequate model exists to specify the other coefficients. The cosine asymmetry factor [equation (5.7b)] is given by £ = ~~ \iobfios2 g sin gdg = - bx / 3 ; hence, bx=-H.
(6.6)
An additional constraint is that p(g) cannot be negative anywhere. For example, if it is desired to represent the phase function by a first-order expansion, p(g) = 1 + ^ c o s g , then bx is restricted to the range — l
108
6. Single-particle scattering: irregular particles
particles are double-lobed, which requires at least a second-order expansion: p(g) = 1 + b^ig) + b2P2(g). 6.D.2. The Henyey-Greenstein function
Henyey and Greenstein (1941) introduced the empirical phase function (6.7)
where £ is the cosine asymmetry factor, £ =
It is isotropic [p(g) = l] when £ = 0. At g = 0 and TT, pig) has the values ( l - £ ) / ( l + f) 2 and (l + £ ) / ( l - £ ) 2 , respectively. If £ > 0 , pig) increases monotonically between 0 and ir, and decreases monotonically if £ < 0 . The angular width of the peak depends on £, Figure 6.3. Henyey-Greenstein particle phase function for several values of the asymmetry parameter.
P(g)
30
60
90 g (degrees)
120
150
180
6.D. Empirical scattering functions
109
becoming more narrow as |£| increases. In the limiting case, as £ -> 1 (or - 1 ) , p ( g ) - > 0 everywhere except at g = ir (or 0), where it becomes infinite in such a way that the integral equals 4ir. Hence, this function is often used to represent the diffraction peak. The expansion of the Henyey-Greenstein function in Legendre polynomials is (Kattawar, 1975)
y=o
(6.8)
One disadvantage of the function is that it has only one lobe. In order to represent a double-lobed phase function, two HenyeyGreenstein functions of opposite symmetry are required. 6.D.3. The Allen diffraction approximation
It was shown in Chapter 5 that the light scattered by diffraction around a sphere is identical with that diffracted through a circular aperture of the same radius in an opaque screen, and consists of a strong central peak surrounded by a series of weaker fringes. A noncircular opening has a qualitatively similar diffraction pattern, but the main peak and the fringes are not azimuthally symmetric. For instance, the angular size and spacing of the diffraction pattern of a rectangular slit is wider along the direction perpendicular to the long edge of the slit than along the direction perpendicular to the short edge of the slit. Jenkins and White (1950) give a number of examples of the diffraction patterns of openings of various shapes. Ensembles of large, irregular particles have narrow, forwardscattered main diffraction peaks, but the fringes tend to be averaged out, so that the diffracted intensity decreases monotonically from the center of the pattern and approximates the envelope of the diffraction pattern of a sphere. Allen (1946) has pointed out that a good empirical fit to the envelope of the diffraction fringes is provided by the function
where the coefficient of X3 in the denominator comes from normalizing pd(g). This function is mathematically more tractable than equa-
110
6. Single-particle scattering: irregular particles
tion (5.24). The exact and approximate diffraction functions are compared in Figure 6.4. 6.D.4. Lambert and Lommel-Seeliger spheres
Suppose an element of area dA of the surface of a sphere is illuminated by collimated light of irradiance / making an angle I with the normal to dA and observed from a direction making an angle e with the surface normal. The phase angle g is the angle between the directions to the source and detector as seen from the particle. Suppose each element of the surface scatters the incident light into unit solid angle according to a reflectance function
dI =
JY(l,e9g)dA.
Then the nondiffractive scattering efficiency Qs and phase function of an opaque sphere, each of whose surface elements scatters light as described by Y(l, e,g), may be calculated as follows. It is assumed that the particle is sufficiently absorbing that internally transmitted light can be neglected. Choose a spherical coordinate system with the origin at the center of the particle and such that the great circle through the sub-source point and the sub-observer point forms the equator of the sphere, with the central meridian of longitude passing through the sub-observer point. The sub-observer Figure 6.4. Allen approximation to the diffraction function (dashed line) compared with the exact expression (solid line).
0.1
0.01
0.001
10" 10
12
14
16
6.D. Empirical scattering functions
111 point and sub-source point are separated in longitude by an angle equal to the phase angle g. The geometry is shown in Figure 6.5. The element of area dA = a2 cos LdLd A is located on the surface of the sphere at longitude A and latitude L, where a is the radius of the particle. The outward normal to the surface makes an angle I with the incident illumination, and an angle e with the direction to the observer. From the law of cosines for spherical triangles, cos I = cos(A + g)cosL, and cos e = cos A cos L. The radiance scattered from the sphere into the direction toward the observer is the integral of Y(i,e,g) over all areas of the surface that are both visible and illuminated: ir/2-g
/
/-T7-/2
JL-
JY i e
2
( > >z)dA-
(6.10)
Two widely used reflectance functions are Lambert's law, Y( i, e, g) = — cos I cos e,
(6.11)
and the Lommel-Seeliger law, ^r,.
x
1 cos ^ cose cos I 4- cos e Figure 6.5.
(6.12)
112
6. Single-particle scattering: irregular particles
These functions are discussed more fully in Chapter 8. Note that both are independent of g. Inserting (6.11) into (6.10) gives f
1=1 J
rir/2 J t I —7rcos 6 cos e dA A=-Tr/2 L = -ir/2 J
= —a 2f 77
/
/
f
• A=-7r/2- L = -ir/2
cos(A + g)cosAcos 3 LdLdA.
Using the identity cos x cos y = [cos(x + y) + cos(JC - y)]/2, this integral becomes
Both integrals are readily evaluated and give / = Ja2\ \ [sin g + (IT - g)cos g ] .
(6.13)
Now, by definition, I = J
This expression was first derived by Schonberg (1929) and is known as the Schonberg function. It may readily be verified that this phase function satisfies the normalization condition, equation (5.6a). It is plotted in Figure 6.6. Similarly, inserting the Lommel-Seeliger function (6.12) into (6.10) gives ~8 r7"/2 ^ cos(A + g)cosLcosAcosL •'A = - TT/2^L = - TT/2 ^ cos( A + g)cos L + cos A cos L
T—f'
7r/2
J r/2~8 = ~i—a 211 477
/
r/2 I
/
cos(A + g)cosA 2zr . . . . j \ r^ rcos LdLdA.
• A=-^/2- L=-1r/2COS(A + g)+COSA
The integral over L is readily evaluated. The integral over A may be evaluated by putting x = A + g/2, so that cos(A + g) = COS(JC + g/2)
113
6.D. Empirical scattering functions
and cos A = COS(JIC - g/2), giving / 2 r/2~g/1 \s e c 8 c o s / = T^" / o2" 16 ^ = - 7 r / 2 + g / 2 L
x
tan sec x ~ s i n o" o" 2 2 J\dx
Hence,
= ~[l-sin|tan|ln(cot|)]. However, this phase function is not normalized. To normalize p(g), set n
&f
2 [l - sin | tan | ln(cot | )|2TT sin grfg = 4TT,
where J / is the normalization constant. This integral may be evaluated by letting y = cos(g/2), noting that ln[cot(g/4)]=ln[(l+y)/(l-y)], and integrating by parts. This gives the phase function of a LommelSeeliger sphere,
-sin|tan|ln(cot|)].
4(1-In 2)
(6.15)
This function is plotted in Figure 6.6 along with the Lambert sphere. Figure 6.6. Lambert and Lommel-Seeliger particle phase functions. 3
2.5
• Lambert
P(g) 1.5
0.5
30
60
90 g (degrees)
120
150
180
114
6. Single-particle scattering: irregular particles
6.D.5. Miscellaneous functions
Several other semiempirical scattering functions have been introduced in the literature to describe various aspects of scattering by large irregular particles. One of the simplest is that used by Hapke and Nelson (1975) in connection with a study of the clouds of Venus. For the diffracted component of the intensity they used a delta function, and for the nondiffracted portion of the scattering efficiency they assumed a function of the form Qs = Se + (l-Se)e-aD, (6.16) where Se is the integral of the Fresnel reflection coefficient given by (5.30b). They assumed that the nondiffracted part of the phase function is isotropic. For large particles, Pollack and Cuzzi (1980) described diffraction by an approximation to the circular-hole equation, and external surface scattering by the Fresnel equations. They represented the internally refracted light by a simple expression of the form cea~*8\ where o and & are empirical constants. 6.E. Some theoretical models
It was shown in Chapter 5 that when X»1, solutions for the scattering of light from a sphere obtained using geometric optics give reasonably good approximations to those using Mie theory. This is especially true for small scattering angles. This suggests that ray theory may be used on large, nonspherical particles, and this approach has been used by several authors to calculate scattering by particles of regular shape, such as cubes, parallelepipeds, hexagonal cylinders, and other common crystalline forms (Liou and Coleman, 1980; Liou, Cai, and Pollack, 1983; Muinonen et al, 1989). However, ray tracing is obviously impractical when dealing with large numbers of particles without regular shapes. Thus, various other schemes for calculating scattering have been proposed. Further references and discussions can be found in the work of Bohren and Huffman (1983) and Schuerman (1980). Most of the models divide the problem into several parts, typically diffraction, external surface scattering, and internal refraction and scattering, and treat each part separately. Often one part is emphasized, and the rest either ignored or treated in an ad hoc fashion.
6.E. Some theoretical models
115
Leinert et al. (1976) fitted observations of the zodiacal light to models in which the particles are large compared with the wavelength. The diffracted radiation was taken to be identical with that from a sphere [equation (5.24)], and surface-reflected light was described by the Fresnel reflectances [equation (5.26)]. Internally refracted light was assumed simply to add a constant term to the phase function. Chylek, Grams, and Pinnick (1976) pointed out that the large resonances displayed by scattering from a perfect sphere, such as peaks in the scattering efficiency and the cloudbows and glories in the phase function, can be attributed to specific terms in the series expansion of the Mie solution. They set these terms equal to zero and assumed that the resulting expression applies to irregular particles. However, though this assumption may be reasonable for particles whose sizes are comparable to the wavelength, it is certainly not true for large, irregular, nonuniform particles. In the extended-boundary-condition method (EBCM) (Barber and Yeh, 1975) the scattering and absorbing centers inside a particle are replaced by induced currents on its surface. In principle, this method can calculate the wave scattered by a particle of arbitrary size and shape. In practice, the calculations converge too slowly to be practicable for particles larger than a few times the wavelength. Kattawar and Humphreys (1980) built up particles from dipoles and numerically calculated the coherent scattering from such ensembles. They also calculated the scattering from two such particles. This method is very powerful, but is highly computer-intensive. In general, the scattered wave can be considered as the sum of waves radiating from all the different points within the particle. The amplitudes of these wavelets are proportional to the amplitude of the wave inside the particle and the difference between the scatterer and a vacuum. Several workers have used the so-called eikonal approximation, in which the internal wave is replaced by the incident wave phase-shifted along undeviated linear paths. Both Chiappetta (1980) and Perrin and Lamy (1983) used this approach to calculate the intensity forward-scattered into large phase angles, and they used an empirical model to describe the scattering at smaller phase angles. Their models assume that the particles have very rough surfaces. They also assunie that the broad backscattered peaks in the particle phase functions are caused by shadows on the surface. Internally refracted light is mainly ignored.
116
6. Single-particle scattering: irregular particles
Schiffer and Thielheim (1982a, b) used a geometric-optics approach to calculate the effects of shadows on light reflected from a rough surface. Mukai et al. (1982) modeled multiple reflections between facets on the rough surface of a particle using the equation of radiative transfer, although it is not clear that this equation is applicable. Internally refracted light is ignored in the model of Mukai and associates, so that the types of particles to which it may be applicable are extremely limited. Emslie and Aronson (1973) and Aronson and Emslie (1973, 1975) emphasized the importance of surface roughness to the spectral scattering properties of the particle. In an elaborate numerical model, later extended by Egan and Hilgeman (1978), they assumed a particle with a basic spherical shape. The absorption and scattering of this sphere were calculated using geometric optics. Roughness on the particle surface was assumed to scatter like randomly oriented dipoles distributed over the surface of the sphere. It was also assumed that the rest of the surface scattered light diffusely instead of specularly. This model has had considerable success in explaining features in the reflectance and emittance spectra of powders in the thermal infrared.
6.F. Experimental studies of scattering by large particles
Experimental studies of scattering by large, irregular particles have been carried out by several workers, including Richter (1962), ZeruU and Giese (1974), Zerull (1976), Hapke and Wells (1981), Weiss-Wrana (1983), Mustard and Pieters (1989), and McGuire (1993). McGuire's work is especially interesting because she systematically varied shape, surface roughness, absorption coefficient, and density of internal scatterers, and because hers is the only work to investigate the effect of internal scatterers. Based on all of these studies, a number of general inferences concerning the scattering properties of large, irregular particles can be made. These are illustrated by Figure 6.7, which compares the intensity and polarization of an irregular particle with those for a particle of spherical shape. (1) If a particle is clear, that is, if it has few internal scatterers and is not completely opaque, its phase function will be dominated by a strong, broad, forward-scattering lobe caused by refracted, singly transmitted rays. This lobe is weaker for irregular particles than for spherical ones, but it occurs for particles of any shape or degree of
111
6.F. Experimental studies of scattering: large particles
surface roughness. Its amplitude decreases as the absorption coefficient increases. At intermediate angles between the cloudbow and the forward-refracted peak the intensity scattered from smooth-surfaced particles is small and independent of absorption coefficient, and the only rays that contribute to the scattered intensity at these angles are those reflected from the surface. (2) If a particle is clear, smooth, and nearly spherical in shape, its phase function will have narrow peaks in the backscatter direction that can be identified with the glory and cloudbow. However, if the particle is irregular or has a rough surface, the amplitude of the forward-refracted lobe will be decreased, and the glory and cloudbow will be smeared out into a single broad, weak backscattered peak of slightly increased intensity. (3) At intermediate angles between the cloudbow and the forwardrefracted peak in a spherical particle the scattered intensity will be small and independent of absorption coefficient, and the only rays that will contribute to the scattered intensity at these angles will be those reflected from the surface. In an irregular particle the intensity at intermediate angles will be increased markedly, often by an order of magnitude, over that for a sphere and will be sensitive to the absorpFigure 6.7a. Measured phase functions of a large, clear, spherical particle with a smooth surface and the same particle after the surface had been roughened. 0.5
- Glass Sphere
G)
n •
0.4 Smooth
^ ^
0.3
A' J
Rough ^ ^ ^
£1 CD
E
0.2
-
f
0.1
-x 30
60
—x^y *
90 120 g (degrees)
150
180
118
6. Single-particle scattering: irregular particles
tion coefficient. This is one of the places where Mie theory is grossly incorrect for an irregular particle. In an irregular particle, energy evidently is redirected from the strong forward-refracted peak of a sphere and is internally scattered into the sideways and back directions. The phase function will be nearly constant or weakly backscattering at small and intermediate phase angles and will have a broad peak in the forward direction. Although certain regular particles with smooth external surfaces that meet at angles close to 90°, such as cubes, may display sharp peaks near zero phase because of an internal corner-reflector effect (Liou et al., 1983), such effects are not important for particles of irregular shapes. (4) An irregular or rough-surfaced particle will have a higher single-scattering albedo than will a sphere of similar size and dielectric constant. As discussed in Section 6.C.3, the effect of the irregularities is to decrease the mean path length (D) of the rays through the particle, thus decreasing the amount of light absorbed and increasing the single-scattering albedo. In doing so, the rays from the forward lobe are redirected into smaller angles. (5) The light reflected from the surface will be positively polarized. In a clear sphere, the polarization will peak around 90° and then drop to a small positive or negative value because the forward-refracted lobe will be negatively polarized and will dominate the scattering for g > 2# c . As seen in Figure 6.7b, at large phase angles the polarization of the irregular particle will be less than that for the sphere, but still will go to a small negative value, showing that the polarization of the transmitted lobe is decreased, but is not completely random. (6) The presence of internal scatterers will cause major departures from the predictions of Mie theory, as can be seen in Figure 6.8. If internal scatterers are present, less light will be transmitted through the particle, and more will be scattered out the back and sides after traveling only a short distance through the particle. This will increase both the single-scattering albedo and the amplitude of the phase function at small and intermediate angles. As internal scatterers are added, the amplitude of the forward peak will decrease, and the intensity scattered at small and intermediate phase angles will increase, so that the phase function will become nearly isotropic. Further increasing the internal-scattering coefficient will cause the forward-scatter lobe to disappear completely, and the phase function will acquire a strong, broad backscatter lobe. The polarization will
119
6.F. Experimental studies of scattering: large particles Figure 6.7b. Measured polarization functions of the particles shown in Figure 6.7a. Glass Sphere Smooth
Rough
30
60
90 120 g (degrees)
150
180
Figure 6.8a. Measured phase functions of large spherical particles filled with internal scatterers. The terms "low," "intermediate," and "high" refer to the relative internal scattering coefficients. 0.7
i
•
i
•
i
Spheres with Internal Scatterers 0.6
Low 0.5 ~ 0.4 ) c 0
£
/High 0.3 0.2 0.1 30
60
90
g (degrees)
120
150
180
120
6. Single-particle scattering: irregular particles
peak around 150°, implying that the polarization of the internally scattered light is random, so that most of the polarization of the light scattered from the particle will be due to the surface-reflected rays. The major effects of internal scatterers on both the intensity and polarization have not been widely recognized. However, they are of special importance because internal scatterers are abundant both in laboratory samples and in planetary soils. Most natural particles contain microscopic inclusions and bubbles. Usually particles are polycrystalline, and the internal grain boundaries scatter light. The process of grinding inevitably produces cracks and fractures that radiate from the surface of a grain into the interior (Tanashchuk and Gilchuk, 1978; Skorobogatov and Usoskin, 1982). The lunar regolith (and presumably the regoliths of other bodies as well) contains large numbers of agglutinate particles, which are welded agglomerates of smaller grains and voids that can act as internal scatterers. The particle phase functions of most planetary regoliths are backscattering. It is of interest to note that the only known way to make a particle strongly backscattering is to fill it with internal scatterers. A particularly dramatic effect of inclusions may occur if they consist of submicroscopic particles with large imaginary refractive indices. Because such particles are very efficient absorbers, they may increase Figure 6.8b. Measured polarization functions of the particles shown in Figure 6.8a. 30
Spheres with Internal Scatterers 20
"
i\ " I \ i v
10 ~
i
^°*-cf ^ / ^
• /
• Intermediate
ri - - - Low——-—*\
/*"
V -10
. 1 . 1 . 1 .
30
i
60
.
i
.
i
.
i
.
i
.
i
.
i
.
90 120 g (degrees)
i
.
i
.
i
.
150
i
.
i
.
180
6.F. Experimental studies of scattering: large particles
121
the effective absorption coefficients of their host particles by orders of magnitude. This is discussed further in Section 7.E.9. McGuire (1993) fitted the phase functions of the particles of her study by a second-order Legendre polynomial expansion of the form (6.17) p(g) = l + b1P1(g)+b2P2(g), and also by a double Henyey-Greenstein function of the form 1+c 2
1-b2 1-c l-b2 (1-2b cos g + b2f/2 + 2 (l + 26cosg + 6 2 ) 3 / 2 ' (6.18a)
In order to keep the number of parameters the same in both expressions, the width parameter b was required to be the same in both Henyey-Greenstein functions, and b was constrained to be in the range 0 < b < 1. There was no constraint on c, except that p(g) > 0 everywhere. From equation (6.8) the relation between the two sets of coefficients is bt = 3bc, (6.18b) b2 = 5b2. (6.18c) Because of the greater flexibility of the Henyey-Greenstein function, superior fits were obtained with (6.18a). The first term of (6.18a) describes the backscatter lobe, and the second term the forward-scatter lobe. Because both HenyeyGreenstein functions have the same 6, the widths of the lobes are equal, so that c determines the height of the back lobe, relative to the forward lobe. If c > 0, the particle is predominantly backscattering; if c < 0, it is forward-scattering. The parameter b describes the width of both lobes; the larger 6, the narrower and higher the lobes. It can readily be shown that the mean cosine of the phase angle is given by (cos g) = -g = bc, (6.18d) and the mean square of the cosine is
122
6. Single-particle scattering: irregular particles
are less obvious than are those of b and c in the double HenyeyGreenstein function. Hence, only the coefficients for the latter expression will be discussed here. The regions on the 6-versus-c plane occupied by the various types of particles studied by McGuire (1993) are shown in Figure 6.9. Clear, spheroidal particles have b ~ 0.5-0.7 and c ~ - 0.9. Increasing the absorption decreases w and decreases b slightly, but does not change c appreciably. Making the particle more irregular in shape, but still clear, increases w, decreases fc, and increases c; that is, the particle becomes more isotropic and less strongly forward-scattering. Roughening the surface of the particle or making a composite agglomFigure 6.9. Empirical double Henyey-Greenstein parameters for large silicate, resin, and metal particles of varied shapes, absorption coefficients, and conditions of surface roughness, and containing differing densities of internal scatterers. Decreasing w causes a particle to move a short distance in a direction away from either end toward the center of the L-shaped area. Almost any change that is made to a particle that is initially clear, smooth-surfaced, and spherical decreases b and increases c. Double Henyey-Greenstein Parameters
rough-surfaced metal —
/
high density of internal scatterers
j / \ /
+1
:
medium density of internal scatterers
1 smooth-surfaced metal
/ /
agglomerates
-
rough-surfaced dielectric
/
0
- \ / Vy rre 9 ular \
\ /
low density of internal scatterers
/ \ \
-1
i
i
i
smooth, clear. sDherical
i
0.5
b
i
i
i
i
1.0
6.G. Aphenomenological model
123
erate adds internal scatterers and increases c. As more internal scatterers are added, c increases, but b remains in the range 0.15-0.35; that is, the particle becomes more strongly backscattering, while the lobes remain low and wide. Note that almost any change from a perfect, uniform sphere has the effect of increasing w, decreasing b, and increasing c from the values that would be predicted by Mie theory. For a particle with internal scatterers, decreasing the absorption increases both w and c, but does not affect b appreciably. Smooth metallic particles scatter nearly isotropically, with b and c both small; roughening their surfaces decreases w and increases c (the particle becomes more backscattering), but b does not change.
6.G. A phenomenological model 6.G.I. The generalized scattering efficiency
All of the models reviewed in Sections 6.D and 6.E either require extensive numerical evaluation (and thus are inconvenient to use) or else involve empirical parameters that are not connected in any obvious way to the size, shape, and complex refractive index of the particle. In this section, a model for the scattering efficiency of a large irregular particle will be presented. The model is based on experimental data. Like all the other models, the one derived here is approximate, but it is given in closed analytic form, and its parameters can be related to the fundamental properties of the particle. Following the notation in Chapter 5, write Qs = Qd + Qs-> where Qd is the portion of the scattering efficiency associated with diffraction, and Qs is the remainder. The diffraction term will be discussed first.
6.G.2. Diffraction
For a perfect, isolated sphere, Qd = 1. By Babinet's principle (Section 5.D.4.b), the diffraction of a wave around an isolated obstacle is equivalent to that through a hole of identical cross-sectional size and shape in an opaque, infinite screen. Because the total power in the diffraction pattern of the hole is equal to the power passing through the hole, regardless of shape, the diffraction efficiency of the hole is 1.
124
6. Single-particle scattering: irregular particles
Hence, the diffraction efficiency of each particle of an ensemble of isolated, randomly oriented, irregular particles is also 1. As with spheres, if the particles are close together the diffraction must be associated with the spaces between the particles, rather than with the particles themselves, and Qd = 0. For the portion of the phase function associated with diffraction, the Allen function, equation (6.9), may be used. 6.G.3. Large-angle scattering 6.G.3.a. The equivalent-slab model for Qs
The success of the equivalent-slab model in reproducing Qs for a sphere suggests that we attempt to generalize it to irregular particles. Hence, we again write equation (5.43) for Qs, Qs = Se + (l-Se)±^®.
(6.19)
The various quantities in this equation have the same meanings as in Section 5.E.6. That is, Se and 5,- are respectively the surface reflection coefficients for light that is externally and internally incident, and @ is the internal-transmission factor. The first term of (6.19) represents the light externally scattered from the particle surface. The numerator of the second term represents the light transmitted once through the particle. The denominator of the second term represents light multiply internally scattered through the particle. However, the mathematical expressions for calculating the quantities will not necessarily be the same as in equation (5.43). Each of the quantities in (6.19) will be discussed separately. 6.G.3.b. Exterior-surface reflection
The evaluation of Se is simple if the particle is convex, with smooth, randomly oriented surface facets. Van de Hulst (1957) defined a convex particle as one that "when illuminated from any direction has a light side and a dark side separated by a closed curve on the surface. Any small surface area dS is on the dark side when its outward normal makes an angle < 90° with the direction of propagation of the incident light; it is on the light side when this angle is > 90°." A more intuitive description of a convex particle would be to say that it is one that is without any depressions, dimples, or projections on its surface
6.G. Aphenomenological model
125
that can cast shadows on another part of the surface. Convex particles include most simple shapes, such as spheres, ellipsoids, cylinders, and euhedral crystals of most minerals. The normals to the surface facets of an ensemble of randomly oriented, convex particles are distributed isotropically. This distribution is identical with that of the surface elements of a sphere. Therefore, provided each facet is smooth, the light reflected from the surface facets of the ensemble is the same as for a sphere, so that the phase function associated with external reflection is given by the integral of the Fresnel coefficients, equation (5.30b). The effects of surface roughness on the scattering characteristics of the particle depend on the scale of the roughness. The discussion in Section 4.F implies that under most circumstances a particle surface may be treated as smooth if the scale of the surface roughness is small compared with the wavelength. More sophisticated theoretical analyses (e.g., Berreman, 1970) indicate that small asperities and depressions in a surface will scatter light in a manner similar to dipoles suspended just above the surface. Because a dipole scatters light proportionally to (size/A)4, if the Rayleigh criterion is satisfied then the effects of a few small surface imperfections usually can be ignored. This conclusion is supported by the microwave experiments of Zerull and Giese (1974), which showed that a sphere with surface roughness elements small compared with the wavelength scatters light in a manner that is virtually indistinguishable from the predictions of Mie theory. A major exception may occur when the complex refractive index is in the region of anomalous dispersion in the vicinity of a strong absorption band. As discussed in Chapter 5, when Ke = m2 - - 2 , a resonance can occur that may cause the scattering and particularly the absorption efficiencies of small particles to be anomalously large, so that they cannot be ignored. This point has been emphasized by Emslie and Aronson (1973). The effect can be especially important in the thermal infrared, where materials typically have strong restrahlen bands. A second exception occurs when k is large, as with metals. Figure 6.10 shows that roughening the surface of a metal will markedly decrease the scattering efficiency. The probable explanation is that the scratches and corners on the surface act like Rayleigh absorbers (Section 5.D.2).
126
6. Single-particle scattering: irregular particles Figure 6.10a. Measured phase functions of large steel spheres with smooth and rough surfaces. 0.1
•
i
•
i
•
i
•
i
•
Steel Sphere
0.08 Smooth
£
0.06
c
CD
•*—•
•x- •
—
0.04 Rough 0.02
30
60
90
120
150
180
g (degrees)
Figure 6.10b. Polarization functions of the particles shown in Figure 6.10a. 50
40
30
20
10
-10
30
60
90 g (degrees)
120
150
180
6.G. A phenomenological model
127
The case in which the scale of the roughness is larger than the wavelength is more difficult to treat, because shadowing may be important. Several experimental studies of the effect of wellcharacterized, large-scale roughness on the reflection from a surface have been carried out, including those by Torrance and Sparrow (1967) and O'Donnell and Mendez (1987). An important conclusion that may be drawn from these studies is that the general character of the scattering remains quasi-specular. That is, the scattering does not change from specular to diffuse, as has been conjectured by many authors. Rather, the reflected light is redirected into a relatively broad peak that is approximately centered in the specular direction. Even at glancing incidence and reflection, the character of the scattering remains quasi-specular. Apparently, a ground or frosted surface scatters light diffusely because of the subsurface fractures created by the grinding process, rather than because of the irregular surface geometry. Thus, to a first approximation the scattering from the surface of an irregular particle may be treated as specular, and it will be assumed that the probability of an incident ray encountering a surface facet oriented in a given direction is the same as for all other facet orientations. Except in regions of anomalous dispersion, Se will be given by the integral of the Fresnel reflection coefficients, equation (5.30b). Figure 6.11 plots Se for a variety of values of n and k. Figure 6.11. External surface reflection coefficient as a function of the real part of the refractive index for several values of the imaginary part.
0.8
0.6
0.4
0.2
2
3 n
128
6. Single-particle scattering: irregular particles
A convenient empirical approximation to Se, which is accurate for k2 <: 1 and 1.2 < n < 2.2, is n
_^ (n-lf+k2 + 0.05. (n + iy+k2
(6.20)
This expression is plotted, along with the exact values, in Figure 6.12. For particles with very rough surfaces containing large numbers of small scattering elements, it will be assumed that the surface asperities and depressions behave like quasi-independent small scattering particles, so that the effective scattering efficiency can be treated like a mixture of large and small particles. Experimental support for this assumption is given in Section 6.G.6. Scattering by mixtures is treated in Chapter 10. 6.G.3.C. Internal surface reflection
In a perfect sphere, the angle at which a refracted ray is incident on the inside of the sphere is equal to that at which it enters the sphere, so that St = 5 e . However, as emphasized by Melamed (1963), for an ensemble of irregular particles the two angles are uncorrelated, and a refracted ray will have virtually equal probabilities of encountering interior surfaces at all orientations. Thus, to a good approximation, a Figure 6.12. External (Se) and internal (S() surface reflection coefficients versus the refractive index for k
1.5
2.5
n
3.5
6. G. A phenomenological model
129
reasonable expression for the interior reflection coefficient is given by the average of the internal Fresnel reflection coefficients over all angles:
5. = fir/2[R±
(ft') + /?,,(#')]cos tf'sin ft'dft', (6.21)
where ft' is the angle of incidence for the interior rays. This expression is identical in form with (5.30b) for Se, except that now R±ll(ft') refers to internally reflected light. In calculating Si9 only the case where k
cos tf'sin ft'dft' + / ^ c o s #'sin ft'dft' .
(6.22)
Carrying out this integration gives (6-23) This approximation for St is plotted as a function of n in Figure 6.12. 6.G.3.d. The internal-transmission factor
One might try to estimate @ by carrying out ray-tracing calculations on a variety of models of specific shapes, hoping that the results would Figure 6.13. Schematic diagram of the model for the scattering efficiency. AD f
1-f
Exponential
Melamed
Internal Scattering
Double Exponential
130
6. Single-particle scattering: irregular particles
have some resemblance to the transmission of irregular particles. However, this approach is of questionable value, particularly if it is realized that accounting for effects of particle shape is not the only difficulty: Often the particles of interest are not clear, but are full of internal scatterers. Instead of that approach, we will consider several rather general, intuitive models and use the results of experiments to choose the best one. Four possible models for @ will be considered: exponential, Melamed, internal scattering, and double exponential. These are illustrated schematically in Figure 6.13. The exponential model It was shown in Chapter 5 that for a sphere the internal-transmission factor can be approximated by an exponential function. Hence, the first and simplest model to be considered is of the form 6-«-«<">, (6.24) where (D) is the average distance traveled by all transmitted rays during one traverse of the particle. If the particle is spherical, (D) 0.9D. However, if the particle is irregular, then (D) can be quite different from D and in general will be smaller. The Melamed model Another possibility for @ was suggested by Melamed (1963). In his model, the external reflection coefficient is assumed to be Se, and Melamed was the first to introduce the idea that the internal reflection coefficient should be 5,-, which we have adopted here. To calculate @, Melamed assumes a clear particle of spherical shape and diameter D = 2a. Radiance emerging from any point of an inner surface after being either transmitted or reflected is assumed to have an angular distribution given by Lambert's law, (1/TT)COS#', where d' is the angle between the surface normal and the direction of the radiance. See Figure 6.14. An incremental area dA a distance x = D cos #' away on the inner surface of the sphere subtends a solid angle dAcosfi'/x2 from the point. The radiance is attenuated by a factor e~ax in traveling from the point to dA. Thus, the total fraction of light emitted from the point that reaches all areas on the interior surface of the particle is c*
©= /
1
— cos# e
i
ax
-
x2
6.G. Aphenomenological model where ty = IT —2ft. Now, dA = 2TTa2sin if/di/j = lirxdx.
0
131 Hence,
/V < "§ *
Generalizing to irregular particles by replacing D by <£>) gives 2
(6.25)
The internal-scattering model The importance of internal scatterers in natural particles has already been emphasized. One approach to investigating their effects on @ would be to attempt to find the radiance inside a spherical absorbing particle containing embedded scatterers. In order to do that it would be necessary to solve the equation of radiative transfer in a spherical geometry. The radiativetransfer equation will be introduced in Chapter 7, with approximate solutions obtained in subsequent chapters. However, the solution of this equation is a formidable problem that has been accomplished to a reasonable degree of rigor only for planar geometry. Thus, we seek a simpler approximate analytic model. Figure 6.14. Melamed model for the internal-transmission factor.
132
6. Single-particle scattering: irregular particles
We have seen that the transmission factor of a perfect sphere can be approximated quite well by e~a
where
rt + exp( - ]/a(a _
r,exp( - ja(a
If the slab is clear, with 4 = 0, then © = e~a2^, showing that the mean path length through the slab is 23*. Thus, we replace 22? by (D)9 where (D) is to be interpreted as the length of the average ray that traverses the particle once without being scattered. This gives, for the internal-transmission factor of an absorbing, scattering equivalent slab, l+r I .exp(--/a(a+ The double-exponential model. It was suggested in the discussion of Se that surface asperities might scatter as quasi-independent particles, so
6.G. Aphenomenological model
133
that Qs can be treated as a mixture of particles of two different sizes. Internal scattering elements located just under the surface might also behave in this manner. Thus, we will consider a model of the form ,-aAZ)
(6.28) where (D) is the mean absorption path through the main particle, AD is the mean path associated with scattering by the surface asperities or subsurface fractures, and / is the fraction of light scattered by these small "particles." 6.G.4. Comparison with experiment
The scattering efficiencies corresponding to these four models for @ are illustrated in Figure 6.15. For very small values of a, all models attenuate light exponentially; that is, @^exp( — a(D)). At large values of a{D) the scattering efficiencies all approach Se, but for the last three models they do so more slowly than for the exponential model, so that Qs is larger at intermediate values of a(D). Figure 6.15. Scattering efficiency of a particle of refractive index n = 1.50 versus a(D) for the four internal-transmission models considered in the text. For the internal-scattering model, 4
Double Exponential Internal Scattering Q
Melamed
s
0.1
Exponential
10
12
14
16
18
20
134
6. Single-particle scattering: irregular particles
The exponential and Melamed models have only one free parameter, <£>); the internal-scattering model has two, (D) and 6; and the double-exponential model has three,
1200
by transmittance
E 1000
I 800 600 400 200
0.5
1.0 1.5 Wavelength (micrometers)
2.0
2.5
135
6.G. Aphenomenological model
were measured over the same wavelength range. Using the reflectance theory derived in Chapters 8-11, and assuming that the powder particles scattered approximately isotropically, Qs{\) was found. Because a(A) was known, <25(A) could be converted to Qs(a). The four models for © were then fitted to the data. The resultant values and theoretical curves for each size range are shown in Figure 6.17. (To avoid clutter, curves for all of the models are not shown in all of the figures.) From these figures several conclusions may be drawn: (1) The general expression, equation (6.19), provides a satisfactory description of Qs as a function of a(D) over the range from 0 to 22. The goodness of the fit depends on the choice of @. (2) Because all four models behave exponentially when a(D) is small, they all fit the data for the smallest size fraction, < 37 fim (Figure 6.17a), for which a(D) < 2. (3) For larger size fractions the exponential model is unsatisfactory. The particles are too bright for a(D) > 2.
Figure 6.17a. Scattering efficiency of the cobalt glass of size < 37 jum as a function of absorption coefficient. The points show the values calculated from reflectance measurements. The lines show the fit of the various models discussed in the text. The Melamed model with (D) = 24 fim and the double-exponential model with (D) = 18 /xm, / = 0.2, and AD = 6.0 /xm are indistinguishable from the internalscattering curve. <37|im _ 0.8
0.6 Internal
Q
Scattering
0.4
0.2 Exponential,
0.1
0.05 a (urn"1)
0.15
136
6. Single-particle scattering: irregular particles Figure 6.17b. Same as Figure 6.17a for size 37-49 jum. The internalscattering curve is indistinguishable from the Melamed model with
0.6 Q
Internal
s
Scattering
i, s=.06jim" 0.4
0.2 Exponential,
0.1
0.05
0.15
Figure 6.17c. Same as Figure 6.17a for size 37-74 /xm. The internalscattering curve is indistinguishable from the double-exponential model with (D) = 45 /xm, / = 0.2, and AD = 4.5 /xm. 37-74 0.8
0.6
0.4
0.2 . Exponential,
0.05
0.1 a (urn"1)
0.15
137
6.G. A phenomenological model Figure 6.17d. Same as Figure 6.17a for size 74-150 fim. The internalscattering curve is indistinguishable from the double-exponential model with
Q
jim _
/Internal Scatterers,
s
0.4 / Melamed,
0.2 Exponential,
0.1
0.05
0.15
Figure 6.17e. Same as Figure 6.17a for size > 150 /xm. The exponential and Melamed models fall well below the measured points. This is the only example for which the internal-scattering and double-exponential models are noticeably different.
Q
s Internal
Scatterers 1
s=.06um" ,
x
Double Exponential , f=.15,
0.1
0.05 a (urn* )
0.15
138
6. Single-particle scattering: irregular particles
(4) The Melamed model is an improvement over the exponential model, but is also inadequate for the larger size fractions. (5) Both the internal-scattering and double-exponential models provide satisfactory fits over the entire range of particle sizes and absorption coefficients measured. The double-exponential model gives a slightly better fit for the largest particles (Figure 6.17e), but at the cost of requiring an additional free parameter. (6) The mean ray path length (D) is of the same order of magnitude as the particle size, but is smaller than the average particle diameter, as is expected for irregular particles. (7) The excellent fit provided by the double-exponential model supports the hypothesis that surface asperities and subsurface scatterers can be treated as separate small particles in their effects on Qs. For the particles of ground silicate glass whose scattering efficiencies are shown in Figure 6.16, the surface imperfections evidently scatter about 15% of the light and behave as particles whose effective sizes are approximately 5 jam. (8) The internal scattering coefficient is 4 = 600 cm" 1 for these particles. It is likely that both of the successful models, the double-exponential and internal-scattering models, are describing the same phenomenon, the scattering by surface asperities and subsurface fractures, but in mathematically different ways. Hence, the decision as to which one to use is somewhat arbitrary. We will adopt the internal-scattering model, equation (6.27), because this model contains the minimum number of free parameters that can still adequately describe Qs. Note that the exponential model is included as a special case when 6 = 0. If it is desired to use the double-exponential model, this can easily be done using the single-exponential model for © and the theory for intimate mixtures that is discussed in Chapter 10. Some additional properties of the internal-scattering model should be noted. When a(D)
Qs~Se+{l-Se)e-°
6.H. The espat function
139
than the mean diameter of the particle. As a increases, @ is influenced more and more by 6, until, when a(D) » 1 , @ ~ 6/4a, and
In this case, Qs is very nearly independent of (D), and the critical internal path is the scattering length 1/6. That is, the portion of the light that has been refracted into the particle and scattered back out has traveled a mean distance of the order of 1/6. Because a(D) » 1, the particle is nearly opaque, so that most of the refracted light interacts only with a layer of the order of 1/ 6 thick on the side of the particle facing the source. Hence, 6 does not necessarily refer to the scattering coefficient throughout the whole interior of the particle, but characterizes the imperfections close to the surface. For this reason, 6 will be referred to as the near-surface scattering coefficient.
A caveat concerning this model for Qs must be emphasized, in that it is based on measurements on one type of particle, pulverized silicate glass. The extent to which these particles are representative has not been explored. Of particular interest is the appropriate value of 6 for other types of particles.
6.H. The espat function
At this point it is convenient to introduce a quantity called the espat function, which is often very useful under certain limited conditions. When k
Under circumstances in which diffraction may be ignored, Qd = 0, QE = 1. Qs = Qs> and W = (1 - Qs)/Qs. Inserting expression (6.19) for
140
6. Single-particle scattering: irregular particles
Qs into (6.29), W may be written in the form
w=
1-5,.
Using (6.27) for ©, the quantity 1 / 0 - 1 may be expanded to give
Hence, when a(D) <^ 1, £a(D). (6.32) 1-5, In Figure 6.19, W is plotted versus [(1 - Se)/(\ - 5,)]a< D) for n = 1.50 and several values of !<£>>. This figure shows that when s(D) is not too large, W is approximately linearly proportional to a over a much larger range of a(D) than QA. There is no physical reason for this quasi-linear behavior when a(D) is larger than 0.1; it is simply a serendipitous mathematical coincidence. However, as a(D) becomes large, the slope of W decreases until W saturates at W
W=(l-5£)/5e. (6.33) When a(D) is small, Figure 6.19 shows that W approximates, although it is not exactly equal to, a straight line that is proportional to [(1 — S e )/(1 — Si)]a(D). However, the constant of proportionality Figure 6.18. Absorption efficiency calculated using (6.19) and (6.27) with 5 = 0. On the log-log plot the dashed line has unit slope, showing that
QA a a for a(D) < 0.1.
10
Q
A
0.1
0.01
0.001
0.001
0.01
0.1
10
141
6.H. The espat function
and the length of the linear region both depend on 6. When a < D) «: 1, the constant of proportionality is about f, and the quasi-linear region extends out to [(1-S e )/{\-S t )]a{D) =*6, or a
De = &l^-(D),
(6.34b)
and g7 is a constant that depends on 4, but is roughly ^ = 1. Thus, the W function is an absorption optical thickness. Hence, this quantity will be called the effective single-particle absorption-thickness function, or espat function.
Except in regions of anomalous dispersion, n is very nearly independent of wavelength, so that De is approximately constant also. For many substances of interest in remote sensing, De/(D) is somewhat greater than 2, while (D)/D is somewhat less than 1, where D is the particle size. Hence, in the absence of more definite information, a Figure 6.19. Espat function of a particle calculated using (6.19) and (6.27) for several values of 4
w
142
6. Single-particle scattering: irregular particles Figure 6.20. Espat function versus absorption coefficient for four size fractions of the silicate glass whose spectral absorption coefficient is shown in Figure 6.16. W is calculated from the measured bidirectional reflectance of the powdered glass; a is measured from transmission. Also given in each figure is the effective particle size De calculated from the slope of the linear part of the curve, and the average path length (D) and internal scattering coefficient 6 calculated from fitting the internalscattering model to the data. From Hapke and Wells (1981); copyright by the American Geophysical Union.
1000 a (cm"1) 10
2000
37 - 49 ^m
w
= 25 urn s = .06|irrr1
i 500 a (cm"1)
I 1000
i 1500
6.H. The espat function
143
rough estimate of De is
De^2D. (6.34c) The effective particle size De decreases with increasing 4
37 - 74 urn
w = 37 urn s = .06|a.m"1
i
l
500
i
1000
1500-
a (cm"1) 10
74- 150|im
/ D e =146nm v
w
/
—
5= 90(im s
-
- Jr 0
V I
I
I
I
I
I
i
i 1000
500 a (cm"1)
i
i
I 1500
144
6. Single-particle scattering: irregular particles
6.16. The spectral espat function W(A) was calculated from the single-scattering albedos w(A) = <25(A) obtained from the measured reflectance spectra, as described previously. Then a(A) was calculated from W(A) using (6.34a), adjusting De for best fit to the entire spectrum. These data are shown as the points in Figure 6.16. In Figure 6.20, W is plotted against a for four size fractions of the Co glass. The linear regions are prominent in each figure. The values of De calculated from (6.34a) for best fit to the data in the linear region are given in the figures. As a increases, W departs from linearity. Also shown for each data set is the theoretical espat function calculated from the internal-scattering model with the same values of 4 and (D) used in Figure 6.17. Combining relations (6.29) and (6.34a) provides a simple approximation for Qs that is valid within the quasi-linear region:
W
(635)
Because W is not independent of 5
6.1. Summary of the irregular-particle model The radiance scattered by a single particle is given by
6.1. Summary of the irregular-particle model
145
When X « 1 , the dipole approximation may be used to calculate Qs and pig). In the region where X > 1, the irregular particle shapes and the fact that most natural assemblages of particles have a fairly wide size distribution cause the optical parameters to approach their large-Jf values smoothly and monotonically, without the large oscillations seen in monodispersions of spheres. Similarly, resonances such as cloudbows and glories do not occur. When X»1, Q sp(g) can be divided into diffractive and nondiffractive components QsP(8) = QsPs(8) + QaPd(8)The scattering coefficient is given by
(6-36)
Qs = Qd + QsHere Qd is the part of Qs associated with diffraction, _ 11 if the particles are isolated, d \ 0 if the particles are close together.
(6-37)
^ * '
The portion of Qs not associated with diffraction is Qs = Se + (l-Se)y^&.
(6.39)
Here Se is the surface reflection coefficient for externally incident light and is given by (5.30b); an approximate expression is
(„-!)*+**
e
(n + lf+k2
V
7
St is the surface reflection coefficient for internally incident light. If the particle is smooth and very nearly spherical, St = Se. However, if the particle is irregular, St is given by (6.21), which may be approximated by A
n(n The internal-transmission factor is @=
'
^
v
v
I -
(6-41)
'
>
(6 42)
where
^i-lT 1 ^^*
( 6 - 43 )
146
6. Single-particle scattering: irregular particles
(D) is the mean length of rays that traverse the particle once without being scattered, and in general is somewhat smaller than the particle size, and 6 is the internal near-surface scattering coefficient. For pulverized silicate glass, 6 = 600 cm"1, although whether this is representative of other particles is not known. A useful approximation, which may be used if the particle is not too absorbing and if the near-surface internal scattering density is not too large, is
where De is an effective particle diameter of the order of twice the actual particle size. If the particle surface is extremely rough, the surface asperities behave like small, quasi-independent particles, and their effects can be included by treating the particle as an intimate mixture of large and small particles. Mixtures are treated in Chapter 11. If the particle is an agglomerate, the effects of the internal voids and surfaces may be described by 6. The diffractive phase function associated with Qd may be approximated by the Allen function
The phase function p5(g) associated with Qs may be described by a double Henyey-Greenstein function, ( \=1 }
+ c 2
l bl
~ (l-2b cos g + b2f2+
1-c 2
1-b2 (l + 2fccosg + & 2 ) 3/2 ' (6.46)
with the appropriate choice for the coefficients b and c dictated by the nature of the particle, as indicated by Figure 6.9. A clear spheroidal particle has b ~ 0.4-0.7 and c 0.9; a particle that is rough-surfaced or irregular or has internal scatterers has b ~ 0.2 and c ~ - 0 . 9 - + 1.
7 Propagation in a nonuniform medium: the equation of radiative transfer 7.A. Introduction
Virtually every natural and artificial material encountered in our environment is optically nonuniform on scales appreciably larger than molecular. The atmosphere is a mixture of several gases, submicroscopic aerosol particles of varying composition, and larger cloud particles. Sands and soils typically consist of many different kinds and sizes of mineral particles separated by air or water. Living things are made of cells, which themselves are internally inhomogeneous and are organized into larger structures, such as leaves, skin, or hair. Paint consists of white scatterers, typically TiO 2 particles, held together by a binder containing the dye that gives the material its color. These examples show that if we wish to interpret the electromagnetic radiation that reaches us from our surroundings quantitatively, it is necessary to consider the propagation of light through nonuniform media. Except in a few artificially simple cases, the exact solution of this class of problems is not possible today, even with the help of modern high-speed computers. Hence, we must resort to approximate methods whose underlying assumptions and degrees of validity must be judged by the accuracy with which they describe and predict observations. 7.B. Effective-medium theories
One such type of approximation is known as an effective-medium theory, which attempts to describe the electromagnetic behavior of a 147
148
7. Propagation in a nonuniform medium
geometrically complex medium by a uniform dielectric constant that is a weighted average of the dielectric constants of all the constituents. The various approaches differ chiefly in their specifications of the weighting factors. Examples are the models of Maxwell-Garnett (1904), Bruggeman (1935), Stroud and Pan (1978), and Niklasson, Granqvist, and Henderi (1981); see also Bohren and Huffman (1983) and Bohren (1986). One of the most widely used is the Maxwell-Garnett model, which will be summarized briefly. For simplicity, it will be assumed initially that the medium consists of a vacuum of dielectric constant Kel = 1, in which approximately spherical particles of radius a and complex dielectric constant Ke2 are suspended. Both the particles and their separations are assumed to be small compared with the wavelength. Radiation having an electric field of strength Ee is incident on the medium and induces a dipole moment p e in each particle that radiates coherently with the incident light. Consider an element of volume in the medium a small fraction of a wavelength in size. From equation (2.26) and subsequent equations the electric displacement in that volume element is given by De = se0KeEe = ee0Ee+Pe,
(7.1)
where Ee is the external field, Ke is the average dielectric constant, Pe is the electric polarization or dipole moment per unit volume, given by Ve = Afpe, N is the number of particles per unit volume, pe is their induced dipole moment, pe = aeee0Eloc, where ae is the electric polarizability, and E loc is the local electric field. At any particle, E loc is the sum of the applied field plus the fields due to the surrounding particles. To calculate P6, the Clausius-Mossotti/Lorentz-Lorenz model, which was originally derived to describe intermolecular fields, is also assumed to be a valid description of the fields between the macroscopic particles. Then, from equation (2.35), Fe = Naeee0Ee/(l-Nae/3).
(7.2)
The dipole moment of a sphere is given by equation (6.2), p e - 4Tra*ee0Eloc(Ke2 - l)/(Ke2
+2).
As discussed in Chapter 6, this expression also results from the first term in the Mie solution for scattering of a plane wave by a sphere.
7.B. Effective-medium theories
149
ae = 4ira\Ke2-l)/(Ke2+2).
(7.3)
Hence, Combining equations (7.2) and (7.3) gives p =
_
N4Tra3[(Ke2-l)/(Ke2+2)]
[ and inserting this into (7.1) we obtain 3<j>[(Ke2-l)/(Ke2+2)] K 1+ <- l-
' - * ' 1 + l-
•
Equation (7.5) is the Maxwell-Garnett effective-medium expression. Unfortunately, effective-medium theories are not capable of explaining an observation familiar to every child the world over: the color and brightness of the clear sky. The reason is that scattering by the individual particles is neglected. This neglect is often justified by the following argument: Suppose the radiation is observed exactly from the direction into which it is propagating. Then the scattered radiation combines coherently with the incident light to produce a wave characterized by a modified dielectric constant. If the medium is observed at some other angle, all the particles in any increment of volume small compared with the wavelength radiate toward the observer. However, a second volume along the line of sight can always be found whose distance differs from that of the first volume by exactly one-half wavelength. The particles in the second volume also radiate toward the observer, but the contributions of the two volume elements are exactly out of phase and cancel. The fallacy in this argument is that it is valid only if the density of particles is perfectly uniform. However, if the various volume elements do not contain equal numbers of particles, they will not radiate equally, and the coherent cancellation will be incomplete. Let the average number of particles in a volume element be (N), and the
150
7. Propagation in a nonuniform medium
actual number of particles in the / t h volume element be N. =
( W ^
E 57V,) = E W ) 2 ,
(7.6)
where the sum is taken over all the volume elements along the line of sight. The sum over SN^ is zero, because, by definition, the average value of 8Nt is zero. Although the summation in equation (7.6) is over the entire line of sight, the critical distance is the wavelength of light. The medium may be conceptually divided into compartments whose dimensions are A / 2 . Then, if the medium is perfectly uniform, the light scattered by the particles in one compartment will be exactly canceled by light scattered in a close neighbor along the line of sight. However, if the medium is inhomogeneous, the cancellation will be incomplete. The variations in particle density are equivalent to variations in the local dielectric constant or refractive index. Thus, the properties of the medium are effectively controlled by the root-mean-square (rms) average of the fluctuations in the local index of refraction over a volume element whose dimensions are of the order of the wavelength. In gaseous media the density fluctuations can be described by a Poisson distribution in which the mean-square deviation is
((8N,f)-Nf,
(7.7)
so that, in this case, the scattered intensity is proportional to the number of particles along the line of sight. Because the scattering efficiency of a small particle is inversely proportional to the fourth power of the wavelength, blue light is scattered more efficiently than red, thus accounting for the color of the sky. These concepts also have applications to the study of gases near their critical points, where they are known as theories of critical opalescence (e.g., Kocinski and Wojtczak, 1978). However, they are deficient on two counts: They fail to describe how the scattered
7.C. The equation of radiative transfer
151
intensity is modified by further scattering as it propagates through the medium, and they break down completely when the particles are larger than the wavelength.
7.C. The equation of radiative transfer
The formalism that is commonly used to calculate how the intensity of an electromagnetic wave is changed by the processes of emission, absorption, and scattering as the wave propagates through a complex medium is a form of the transport equation known as the equation of radiative transfer. The fundamental assumption of this formalism is that the inhomogeneities of the medium emit and scatter the radiation independently of each other and incoherently. This will be the case when the material consists of discrete molecules or particles that are randomly positioned and oriented. Thus, with care, the theory is applicable to the two media that are of greatest interest in remote sensing: atmospheres and planetary regoliths. However, it will not give correct results if the particles are uniformly spaced and regular in shape. The derivation of the equation of radiative transfer is as follows: Consider a radiance field /U,ft), which describes the intensity of electromagnetic radiation at a point s, propagating into a direction ft, that has been emitted or scattered at least once within a particulate medium. The units of /(s,ft) are power per unit area per unit solid angle. In general, unless stated otherwise, /(s,ft) will also be a function of wavelength or frequency; however, in the interest of economy of notation, this dependence usually will not be denoted explicitly. Suppose s lies on the base of a right cylinder of area dA, length ds, and volume dsdA, where ds points in the direction of ft (Figure 7.1). Then the radiant power at s passing through the base contained in a cone of solid angle dCi about ft is I(s,Cl)dAdCi. Similarly, the power emerging from the top of the cylindrical volume into rfft is I(s + ds,ft)dAdn
= \l(s,ft) + a / ^ a ) ds\ dAdCl.
The difference between the power emerging from the top and that entering at the bottom is (dI/ds)dsdAd£l. More generally, 31/ds is
152
7. Propagation in a nonuniform medium
the divergence of / (see Appendix A) in the direction of ft,
The change in radiant power is due to various processes occurring in the cylinder that add or subtract energy from the beam. In this book we will be concerned with three such processes: absorption, scattering, and emission (Figure 7.2). Those effects of absorption and scattering that subtract energy from the radiance are usually lumped together into a single process called extinction. Each of these processes will now be considered. Extinction. The volume extinction coefficient E(s, ft) of the medium is defined such that the decrease in power APE due to extinction as the beam propagates through the volume element is APE= - E(s,£l)I(s,Q,)dsdAdCl.
(7.8)
If extinction were the only process acting on the intensity, then -dldAdCl, Figure 7.1.
Ks,Q)
7.C. The equation of radiative transfer
153
and the fractional change in the average intensity would be
dI/I=-Eds, which may be integrated to
[ T ]
(7.9)
The volume extinction coefficient may be separated into the volume absorption coefficient K(s,£l) and the volume scattering coefficient S(s, ft), where E(s,n) = K(s,Cl) + S(s9n).
(7.10)
Scattering. Scattering increases as well as decreases the power in the beam, because light of intensity I(s, ft') propagating through the volume element dsdA in a direction ft' can be scattered into the direction ft. The volume angular scattering coefficient G(s,ft',ft) of the medium is defined such that the power APS passing through dsdA into a cone of solid angle dQ! about direction ft' that is scattered into a cone rfft about direction ft is
Figure 7.2. Schematic diagram of the changes in the radiance as it traverses the cylindrical volume dsdA.
154
7. Propagation in a nonuniform medium
To find the total amount of power LPS scattered into the beam, the contribution of intensities traveling through the volume in all directions ft' must be added together, so that A P 5 = f dPsdnf = dsdAdCl-l- f 7(s,ft')G(s,ft',ft)dft'. J
4
4 7 7
J
(7.11)
4ir
The integral of G/4TT over all directions is the volume scattering coefficient, S(s,Cl) = -i-[
G(s,ft',ft)
(7.12a)
For an ensemble of randomly oriented particles, G depends only on the scattering angle 0' between ft' and ft, rather than on the two directions separately. In that case, 5 is independent of ft, and dft' = 2v sin 0' d0f, so that
5(^) = 4 rG(s,ef)sm0'd0f.
(7.12b)
Emission. Let the volume emission coefficient F(s,£l) be the power emitted per unit volume by the element at position s into unit solid angle about direction ft. Then F(s,£l) is defined so that the contribution of the emitted radiation to the change in power is APF = F(s,n)dsdAdn. (7.13) In general, there are at least four processes that may contribute to the volume emission: single scattering, thermal emission, fluorescence and luminescence, and stimulated emission. Fluorescence and luminescence will not be considered further in this book. Stimulated emission is important in gaseous media that are not in thermodynamic equilibrium, but usually not in particulate media. Hence, only single scattering and thermal emission will be considered in detail; their volume coefficients will be denoted by F 5 (s,ft) and FT(s,£l), respectively. Then F(s,ft) = F 5 (s,ft) + F r (s,ft). (7.14a) Single scattering will be discussed first. In a large number of problems the medium is illuminated by highly collimated radiation incident on the top surface from a source that may be considered to be effectively at an infinite distance above the medium in a direction ft0 making an angle I from the zenith. Let / be the irradiance at the upper surface of the medium from such a source. Then at any altitude
7.C. The equation of radiative transfer
155
below the surface of the medium the irradiance that has not been extinguished is, from equation (7.9), J8(£l - a o ) e x p [ - f"E(s',£i) ds'l where 5 is the delta function. The irradiance that is scattered by the medium may be considered to be a source of diffuse radiance that contributes an amount
s,no,fi)exp ~J°°£(s')
(7.14b)
to the volume emission coefficient. The second emission process that we will be concerned with is thermal radiation FT(s,£l). The effects of this process will be considered in detail in Chapter 13, and further discussion will be deferred to that chapter. Equating the sum of all the contributions (AP 5 + &PE + APF) to (8I/ds)dsdAdCl and dividing by dsdAdfl gives
[
(
)
(
)
F ( s , n ) . (7.15)
Equation (7.15) is the general form of the equation of radiative transfer. In most applications the medium is horizontally stratified. Let the positive z axis point in the vertical direction, and let ds make an angle # with dz, so that ds — dz /cos & (Figure 7.1). Making this substitution and dividing through by E(z) gives E(z)
dz
/
~
s(z) i
( ' ) £(z)4ff/lir/^11) z fl
+
S(z)
, F(z,n) E(z)
•
Let r be the optical depth,
r= rE(s')ds'/cos <&= rE(z')dz', J
s
J
(7.16a)
z
so that = -E(z)dz.
(7.16b)
156
7. Propagation in a nonuniform medium
The optical depth T is a dimensionless vertical distance expressed in units of the extinction length \/E. The altitude z can be expressed equivalently in terms of T. Radiance emitted vertically upward at altitude z within the scattering medium is reduced by a factor e~T by extinction as it propagates to the top of the medium. Conversely, light incident vertically on the top of the medium will be reduced in intensity by the same factor e~r as it penetrates to the altitude corresponding to T. Let w(z) be the volume single-scattering albedo, w(z) = S(z)/E(z), (7.17) and let p(z,Cl',£l) be the volume phase function,
p(z,a\n) = G(z,n',a)/s(z).
(7.18)
Define the source function,
n0,n)+FT(T,n),
(7.19)
where i is the angle between the incident irradiance and the vertical, and 9rT = FT/E. (7.20) Then the equation of radiative transfer may be written
4
(7.21) Note that none of the quantities E, S, G, <£, or F appear explicitly in the radiative-transfer equation when it is written in this form, only their ratios in the form of w, p , and &T. The advantage of writing the equation in the form of (7.21) is that in many applications the parameters S, G, E, and FT all have the same dependence on altitude through being proportional to a common function of z, such as the density of the medium. In these cases the ratios w, p , and &T are independent of z or T, which simplifies the problem considerably. In most (but by no means all!) of the applications of interest, K, 5, and E are independent of the direction in which the radiance is
7.D. Radiative transfer with well-separated scatterers
157
propagating. Hence, in what follows it will be assumed that these coefficients are independent of ft, unless explicitly stated otherwise. 7.D. Radiative transfer in a particulate medium of well-separated scatterers
If the medium is uniform, the volume absorption coefficient is identical with the ordinary absorption coefficient: K = a = 4irk/A. The extinction and scattering coefficients then have physical meanings similar to a: The intensity of a collimated beam of radiation penetrating through the material will be reduced by a factor of e" 1 in a distance l/K = l/a if only absorption is present, or in a distance 1/5 if only scattering occurs, or in a distance \/E = 1/(5 + K) if there is a combination of absorption and scattering. The reduction in intensity is uniform across the wave front, and K and 5 are simply constants of the medium that must be specified. However, these definitions clearly are inadequate if the medium is nonuniform and, in particular, if it consists of particles that are large compared with the wavelength. The usual approach in this case is the following: Suppose the medium consists of identical particles separated by distances that are random, but that are, on the average, large compared with both the particle sizes and the wavelength. An example would be a cloud. Consider a slab of the medium of area dA and thickness ds, through which light of intensity / penetrates (Figure 7.3). The volume of the slab dsdA is assumed to be large compared with the volume of an individual particle in such a way that dA is much larger than the geometric cross-sectional area of an individual particle, but ds is so small that no particle shields another within the incremental volume element. Let N be the number of particles per unit volume, a be the particle geometric cross-sectional area, and Q E be their extinction efficiencies, as defined in Chapters 5 and 6. Because the spacing between the particles is large, the efficiencies may be assumed to be the same as if the particles were isolated. There is a total of NdsdA particles in the slab, so that as the light travels through it the intensity intercepted by the particles and extinguished is APE = dJdtl dsdA = JN
158
7. Propagation in a nonuniform medium
ume extinction coefficient for a particulate medium may be defined as E = NaQE.
(7.22)
More generally, if the slab contains a mixture of different types of particles with Nt particles of type / per unit volume, geometric cross section ov, and extinction efficiency QE., then the volume extinction coefficient is defined as (7.23a) where Af ^ is the total number of particles per unit volume, and (VQE^ - ( S ^ o ^ Q ^ p / N is the average extinction cross section. In general, all of the quantities in E may be functions of position, although this is not shown explicitly. Similarly, define (7.23b) and
K-XfN,*,QA/t
(7.23c)
where Qs^ and QA^ are, respectively, the scattering and absorption efficiencies of the /th type of particle. The quantity that describes the angular deviation of the light scattered by a particle is the phase function p.(6) [or p/g)] of the /th type of particle. Hence, the angular scattering coefficient is defined as <X}N,a,Qs,p.(0). Figure 7.3. Extinction by particles in a slab.
(7.23d)
7.E. Radiative transfer with arbitrary separations
159
The expression analogous to equation (7.9) for the transmission of a slab of particles of thickness s is r
c
i
(7.24)
Note that defining the volume coefficients in this way results in a subtle, but major, change in the physical interpretation of the propagation of radiation through the medium. In the uniform medium the absorbers and scatterers are distributed. That is, the radiation encounters them everywhere, so that the intensity is reduced equally across the wave front. However, in the nonuniform medium, the absorbers and scatterers are localized, so that the local intensity is drastically perturbed. Part of the wave front traverses the slab unchanged, but the parts that encounter particles are extinguished. Even if the particles are smaller than the wavelength, the energy is no longer distributed uniformly across the wave front of the transmitted light. It is the intensity averaged over an area of the wave front whose dimensions are much larger than both the particle separation and the wavelength that must be considered to be fractionally reduced by an amount N(aQE) ds in distance ds. Thus, if the equation of radiative transfer is used to treat a medium of discrete scatterers, this formalism explicitly applies only to volumes whose dimensions are large compared with the inhomogeneities in the medium. If the medium consists of particles, the scattering and absorbing properties must be averaged over distances greater than the particle sizes and separations. Because in a gas the fluctuations in the molecular density are equal to the number of molecules per unit volume, and because the fluctuations are uncorrelated, these definitions may be used in equation (7.21) to calculate radiative transfer in atmospheres, as well as in clouds of well-separated particles. However, in order to apply these concepts to media of closely spaced particles, they require certain modifications, to which we now turn. 7.E. Radiative transfer in a medium with arbitrary particle separations 7.E.I. Particles smaller than the wavelength
Suppose we are dealing with a particulate medium in which the particle separation is not large. This case includes all planetary soils
160
7. Propagation in a nonuniform medium
and powders in the laboratory. The question, then, is this: What properties of the medium should be used for the parameters in the radiative-transfer equation? The answer depends on the size parameter X of the soil particles. It was shown in Section 7.B that the scattering properties of a medium are effectively determined by the rms average of the fluctuations in the refractive index on a scale of the order of a wavelength. In a gas, this quantity is proportional to the mean density, so that the expressions developed in Section 7.C can be used for molecules in atmospheres, as well as for particles in clouds. However, many soils, clays in particular, consist of submicroscopic particles in contact. Intermolecular surface forces cause them to be highly cohesive, so that they aggregate into clumps. Hence, their density distributions do not obey Poisson statistics. Ishimaru and Kuga (1982) studied extinction by colloidal suspensions. For X
Nka3 a NaXa3 and S = N
that S' <xaS/K<xl/a\. On the other hand, if X « 1 , it would be 3 expected that Ka Na k/Xa Na3a and SaNa6/X4; hence, S' a a3\~\ Kortum reported that when X » 1, S f a 1/flA, as predicted. However, when X
7.E. Radiative transfer with arbitrary separations inhomogeneities that behave as particles that are intermediate in size between the two foregoing extremes; that is, the size parameter X of the effective particles of the medium is ~ 1. These results support the arguments of Section 7.B: The scattered radiance apparently tends to sample inhomogeneities in the medium that are of the order of the wavelength. This suggests that the following procedure might be appropriate. An effective-medium theory, such as the Maxwell-Garnett model discussed in Section 7.B, can be used to calculate the average refractive index of the medium. This average refractive index can then be used in the expressions developed in Chapters 5 or 6 to calculate the efficiencies and phase function of a particle with X ~ 1, which can then be used in the radiative-transfer equation. This problem is further complicated because interparticle cohesion, which is an attractive force that acts between the surfaces of small particles, causes the medium to separate naturally into clumps that often are larger than the wavelength. If this occurs, it may be that the effective particles should be taken to be the large aggregates, rather than wavelength-sized clumps. Although more experimental work is needed to refine and better quantify Kortum's results and to test these suggestions, it appears that solutions of the radiative-transfer equation can be applied to a closepacked medium of particles smaller than the wavelength, provided that the effective scatters are interpreted as aggregates, rather than the actual particles that make up the medium. The appropriate size of the effective particles is unclear, but probably is of the order of the wavelength; alternatively, it may be of the order of the sizes of the clumps in the medium, and thus may be large compared with the wavelength. 7.E.2. Particles large compared with the wavelength
When the distance between particles is sufficiently great, they can be treated as scattering like isolated single particles, and the definitions of the scattering parameters in Section 7.D are valid. It is of interest to inquire whether these definitions can be suitably modified to be applicable to a medium in which large particles are close together. When X»1 the nonuniformities in the refractive index occur primarily at the particle surfaces, where the refractive index changes over a short distance from that of the particle to that of the material
161
162
7. Propagation in a nonuniform medium
in which the particles are embedded, usually air or vacuum. The region where X » 1 is that of geometric optics, in which the concept of energy propagating along localized ray paths within the medium is a valid approximation. The individual rays cannot be considered as interacting with the entire particle, but only with a small portion of its volume, so that a large number of rays will be required to statistically sample the entire particle. The experiments by Kortum (1969) for X » 1 , discussed in Section 7.E.1, support the idea that the interactions of an ensemble of rays with an Ensemble of closely spaced particles are similar to the interaction of a plane wave with an isolated particle, which was treated in Chapters 5 and 6. However, there are several differences between the two cases that significantly alter the nature of the scattering. These include the role of diffraction, the dependence of the radiative-transfer coefficients on particle density or porosity, shadowing, nonuniform illumination of particles when closepacked in a powder, optical coupling between particles, and coherent effects. Each of these effects will be discussed in the following sections. 7.E.3. Diffraction One major change that occurs as the separation between large particles is decreased was discussed in Chapter 5. When the filling factor
7.E. Radiative transfer with arbitrary separations
163
An upper limit to the mean particle separation for which diffraction can be ignored may be estimated as follows: The major contribution to the diffracted light may be considered to come from rays that pass through an annular zone between a and v^2 a from the center of a particle of mean radius a; the half-width of the cone of diffracted light is given approximately by A0 = A / 4 a (Chapter 5). Thus, a second particle located within a distance Sf -]/2a/kd = 4 / 2 a 2 / A behind a first particle is illuminated by light diffracted by the first particle, but that radiation is not attenuated appreciably by spreading and cannot be distinguished from incident irradiance that has passed unscattered through the diffraction annulus. However, a third particle located at a distance much larger than 3* behind the first is illuminated by light that has been well spread out into its diffraction pattern, and hence the radiation can be interpreted as having been scattered. Of the light scattered into the diffraction cone by a particle in a medium, a fraction e~N(T^ that has not been intercepted within a distance 3* of the particle can be considered as having been diffracted, while the remainder that has been intercepted, 1 - e~N
(7.25)
This condition is certainly met for media consisting of large particles (X^>1) approximately in contact (<£ ~ 0.5). If X is not large, the diffracted light cannot be separated from the rest of the scattering pattern, and the whole discussion is irrelevant. 7.E.4. Thefillingfactor It may easily be demonstrated that (7.22)-(7.24) are incorrect for a medium of large particles when > is not small. Suppose a uniform medium consists of only one type of particle, with constant N, and where a is the write JQEds' = Es in the form Es = NaaQE(s/a), average radius of a particle, a = y/a/ir. The volume of a particle is approximately v- = jira3 = \cra, and the filling factor is (f) = No- = ^
164
7. Propagation in a nonuniform medium
so that t(s) = exp[-%
7.E. Radiative transfer with arbitrary separations
165
without encountering a particle. This is equivalent to the probability that the center of any particle / in the slab is not in a circular cylinder axial to the ray with cross-sectional area cr^QE^ and volume (j'O Sz sec ft In order to calculate this probability, imagine that all the particles are removed from the slab and then replaced at random positions one at a time. For economy of notation, it will be assumed temporarily that QE^ = 1; this restriction will be removed later. Then if the first particle is placed in the slab, the total volume available in which its center can be placed is *V - o-v Hence, the probability that its center is not in the cylindrical volume ax Sz sec ft is atSz sec ft The second particle is now placed in the slab. Because the slab already contains one particle with volume v-v and because the two particles cannot overlap, only a volume "V - v-x - o-2 is available to the center of the second. Thus, the probability that its center is not in the cylindrical volume a2 Sz sec ft about the ray is T' - o-x - o2 - a2 Sz sec 1
ft
2
a2 Sz sec ft 1
2
Continuing the process, the probability that a third particle placed randomly in the slab will not block the ray is o-3 Sz sec ft A = 1 *3 - y_^ + + y In general, the probability that the / t h particle will not block the ray is (7; Sz sec ft A
+ —1_
—t
The probability that none of the JV particles will block the ray is the product of the individual probabilities. Thus, the transmission of the slab for this particular sequence of drawing the particles is
v
a: Sz sec ft \
It is assumed that there is some void space within the slab, because otherwise it would be completely opaque, and there would be no point
166 7. Propagation in a nonuniform medium to this calculation. Then
Because the loading of the slab can be done in any order, to calculate AT completely, all Jf\ permutations and combinations of the order in which the particles are placed must be calculated and the average taken. It is convenient to carry out this averaging beginning with the last ( / = Jf) term. This term is of the form
For each value of n this term occurs (Jf -1)! times, because once n has been specified there are (J^ -1)! ways to choose the earlier terms in (7.28). Thus, the average of the / = JV term in (7.28) carried out in this manner is ** 8z
(a)
8z
where (7.29) is the average particle cross section weighted by number, and (7.30)
is the average particle volume weighted by number. Continuing, the second to the last term ( / = Jf — 1) in (7.28) is of the form 1
v (
\
7.E. Radiative transfer with arbitrary separations
167
For each value of n and k this term occurs (Jf —2)! times. Thus, the average value of the / = Jf — 1 term in (7.28) is
( ^ - 2 ) 1 fiz Because ^f »• 1, negligible error will result if in the last equation o-n is replaced by < o-). Then this expression becomes
Making similar approximations to calculate the average value of the — /)th term gives _8z (a) Tjr e ~ ^ \ The relative error in the sums over the individual volumes in the denominators of each term in (7.28) that results from replacing ^ by < o) becomes progressively larger as I increases. However, for large I these sums involve fewer particles and become progressively smaller compared with 2^. Also, AT is dominated by the terms with small / . Hence, making this substitution in any of the terms will cause negligible error in AT. Thus, to a sufficient approximation, the last expression can be used to calculate the average of each term, giving (73i)
Expanding each fraction in (7.31) and rearranging gives
168
7. Propagation in a nonuniform medium
When JT is large, the approximation L^=1n' - ^ / + 1 / ( / + 0 (Dwight, 1947, equation 29.9) may be used, so that AT becomes
8z(
N(a)8z sec The probability of a ray penetrating a number Jl of layers each of thickness 8z of randomly arranged particles is the product of the probabilities of penetrating the individual layers. Forming this product and passing to the limit of small thickness gives
-> expjsec # It has been assumed that QE^ = 1. If this assumption is not made, then (a) must be replaced by (crQE) in the preceding equation. Now,
secddz = ds; hence,
t(s) = explf^*®^
ln(l - 0) ds'\.
(7.32)
Comparing (7.32) with (7.24) shows that the more general expression for the volume extinction coefficient, which is valid for any particle spacing and size distribution, is (7.33)
By analogy, similar expressions apply to the volume scattering, absorption, and emission coefficients. (Note that the logarithm is negative, so that E > 0.) When
7.E. Radiative transfer with arbitrary separations
169
Most papers on radiative transfer in a particulate medium implicitly assume that (7.23a) is correct. Equation (7.33) shows that the results of these discussions remain valid, provided that in the extinction coefficient the particle density N is replaced by an effective particle density NE given by
M±^>
(734)
7.E.5. Nonuniform particle illumination and shadowing When the particles are large and close together, the following assumptions are reasonable: (1) An ensemble of randomly oriented and positioned particles, randomly illuminated over portions of their surfaces, will absorb and scatter the same average fraction of the incident light with the same average angular pattern that they would if they were uniformly illuminated. (2) The statistical effect of keeping the intensity of rays that penetrate a given distance constant while reducing their number exponentially by extinction is the same as reducing the intensity uniformly and exponentially, insofar as scattering by randomly oriented and positioned particles is concerned. If these assumptions are valid, then Qs and p(g) are unchanged from their values when the particles are isolated, and the expressions developed in Chapters 5 and 6 may be used in the definitions of 5, K, and E, after removing the terms describing diffraction. An important exception to these assumptions occurs when one particle shadows another at small phase angles. This causes a phenomenon known as the opposition effect, which will be addressed in detail in Chapter 8.
7.E.6. Optical coupling between particles When large particles are so close together that they touch, the parts of the surfaces of two particles that are within roughly a wavelength of each other are optically coupled, and the internal and external surface reflections are decreased there. In most cases, even for powder that is closely packed, the fraction of the total particle surface that is this
170
7. Propagation in a nonuniform medium
close to another surface is small, so that in many cases optical coupling can be ignored. However, there are situations where such coupling is not negligible. A major example is a rock. A typical rock consists of a large number of grains of one or more minerals welded together into initimate contact. Under these conditions the relative refractive indices across the grain boundaries are from one mineral to another, not from the mineral to air or vacuum. Consequently, the internal and external surface reflections are greatly reduced, resulting in major changes in Qs and p(g) for each grain in the rock. The general result is that transparent grains are more forward-scattering in the rock than when isolated, and the rock appears darker. This change in coupling is the reason the optical properties of a rock are different from those of a powder made by breaking the rock up into its constituent mineral grains. The decreased contrast in refractive index also explains why adding water to a powder, such as in wet beach sand, or snow with a high moisture content, or suspending it in a transparent matrix, such as a plastic or a varnish, as in paint, makes it darker. For the same reason, subjecting a particulate material to high pressure lowers the reflectance, as shown by Schatz (1966). This effect is especially strong if the pressure is high enough to approach the shear strength of the material, so that the particles deform into close optical contact. They may also go into a more regular packing, thus negating the assumption of random positioning and orientation. However, most particulate media encountered in the laboratory and planetary regoliths are fairly loosely packed, so that the assumption of randomness is likely to be valid in most cases of interest. The radiative-transfer equations will still apply to most media encountered in the laboratory or in the field, but, depending on the circumstances, optical coupling effects may cause the scattering parameters of the particles in the medium to be different than when isolated. 7.E.7. Coherent effects
In the exact Mie solution to scattering from a large sphere, the electric and magnetic fields far from the particle are transverse to the
7.E. Radiative transfer with arbitrary separations
171
direction from the sphere and fall off as 1/ A% where A* is the distance from the center of the particle. However, close to the sphere the fields also contain components that are parallel to the radius vector and that fall off as (A/A*)2. The latter region is called the near-field zone. It is sometimes argued that in a closely packed powder each particle is in the near fields of its neighbors. In addition, the waves are subject to constructive and destructive coherent interference effects as they propagate between the particles. For these reasons it has been questioned that the radiative-transfer equation and the scattering parameters of an isolated particle are applicable to close-packed media. However, when ^ T » l , ( A / ^ ) 2 < c l . Furthermore, the scattering of the light from the large particles of the medium can be described in terms of individual, localized rays interacting with those parts of the particle where changes in the refractive index occur, rather than as a large, plane wave with an entire isolated particle. Thus, if the near-field zones can be considered as falling off as (A/A4)2, then r should be interpreted as the distance to the localized scattering interaction. The near-field zones occupy only a small volume around only parts of each particle, so that only minor portions of neighboring particles will be in them, and near-field effects should be small. It may also reasonably be assumed that when the particles are randomly spaced and oriented, all coherent effects cancel. However, when the particles are of the order of the wavelength in size and separation, a coherent phenomenon known as weak localization or coherent backscatter can occur. This effect is discussed in Chapter 8. 7.E.8. The radiative-transfer coefficients in a medium of large particles with arbitrary separation
In summary, there appear to be ample reasons to believe that the equation of radiative transfer applies to close-packed media and that the scattering parameters of individual particles may be used, after the appropriate modifications discussed in the preceding sections. Applying the discussion of this section to radiative transfer in a medium of arbitrary porosity, the general expressions for the various coefficients that appear in the equation of radiative transfer are given by the following expressions, in which the subscript / denotes a
172
7. Propagation in a nonuniform medium
particular type of particle:
E - NE(aQE) = - Wj+lXjN^QEp S = NE(*QS) = -
H1
K = JV£<(r^> = -
Hl
(7.35)
4*h,Nf*fQSf,
(7.36)
^^N^QAp
(7.37)
N=X;N},
(7.38)
<
(7.39) (7.40) (7.41)
The filling factor is
(7.42)
where
^-(V^J/AT
(7.43)
is the average particle volume, and the porosity is 1 - 4>. Because we are assuming that the individual particles are randomly oriented, the probability of scattering through a given angle depends only on the scattering or phase angle, not on the direction of the incident radiation. Hence,
G(g) = NE(aQsp(g)) = -
H1
^hfNtafQSfpf(g),
(7.44)
where P^(g) is the angular phase function of the / t h type of particle. The generalized definition of the volume single-scattering albedo now becomes w = S/E = (*QS)/(
= (hNnQBi«,)/{h NnQB,),
(7.45a)
where (7.45b) is the single-scattering albedo of the / t h type of particle. Note that the volume single-scattering albedo is just the average of the individual particle single-scattering albedos weighted by their extinction cross-sectional areas. It is convenient to define a quantity called the V,-QS,/QE,
7.E. Radiative transfer with arbitrary separations
173
albedo factor of the medium as y = \/1-H> .
(7.45c)
The volume phase function is
p(g) = G(g)/S = [2,fNf
W = { \ - w)/w = K/S = [
f
M
/
/ W J-
M W f t
( 7 - 47 )
If only one type of particle is present in the medium, then w = w, = p(g)> and W = W. In that case the volume single-scattering albedo and phase function are the same as the single-particle quantities and may be used interchangeably, and if QE = 1, then w = Qs. 7.E.9. Dispersed opaques and scatterers
It was shown in Chapter 6 that internal scatterers can have a major effect on Qs and p(g) of individual particles. Often these scatterers consist of discrete submicroscopic particles dispersed throughout the larger particles, and an important question is how the optical properties of the small inclusions modify those of the large particles. If the inclusions are well separated and randomly positioned, the internal scattering coefficient is given by 5 = NcrQs, where N is the number of inclusions per unit volume, a is their cross section, and Qs is their scattering efficiency, given by equation (5.14), and the internal absorption coefficient is a = at + NaQA, where a{ is the intrinsic absorption coefficient of the large particle, and QA is the absorption efficiency of the inclusions, given by equation (5.13). On the other hand, if the inclusions are closer together than the wavelength, they may interact coherently. In this case their effect on the undeviated ray can be calculated using an altered refractive index, given by the Maxwell-Garnett effective-medium equation (7.5). Unfortunately, as discussed in the first part of this chapter, effective-medium
174
7. Propagation in a nonuniform medium
theories do not take scattering into account. However, they will give accurate results if the inclusions have a large imaginary part of the refractive index, so that they are almost pure absorbers. Because small particles with large imaginary refractive indices are efficient absorbers, dispersed opaques can have a major effect on the single-scattering albedos and reflectances of a particulate medium. This is of considerable importance in the reflectance of the lunar regolith (Hapke et al., 1975) and some shocked meteorites (Fredriksson and Keil, 1963).
7.F. Methods of solution of radiative-transfer problems 7.F.I. Introduction
The equation of radiative transfer is a linear integrodifferential equation. In spite of the fact that it is one of the most important equations in astrophysics and remote sensing, it has proved to be remarkably intractable, and no exact analytic solution in closed form has been obtained. Therefore, numerical computer methods must be used if a high degree of accuracy is desired; otherwise one must be satisfied with approximate analytic solutions. It must be emphasized again that when the radiative-transfer equation is applied to particulate media, the solutions implicitly average over dimensions in the media that are large compared with both the wavelength and the particle separation. In addition to numerical techniques for solving radiative-transfer problems (e.g., Gerstl and Zardecki, 1985a), a large number of different methods have been developed to obtain solutions of varying degrees of accuracy. These are described in many references, including Chandrasekhar (1960), Kourganoff (1963), Hansen and Travis (1974), Sobolev (1975), Ishimaru (1978), Van de Hulst (1980), and Lenoble (1985), and the reader is referred to those works for details of a particular method. In this section I will outline a few that are widely used.
7.F.2. The Monte Carlo method
This numerical method is potentially the most accurate, especially for complicated geometries (including spherical atmospheres) and highly anisotropic particle phase functions, but requires large amounts of
7.F. Solution of radiative-transfer problems computer time: Photons are injected into the medium, and at each computational step they are presented with a probability of being scattered through a given angle or absorbed. Each photon is followed until it either is absorbed or leaves the medium. The process is continued until adequate statistics are built up for all directions of scattering. Photons that contribute to the observed brightness can be calculated as if they travel either from the source to the detector or in the opposite direction, as convenient. 7.F.3. The radiosity method The radiosity (Borel, Gerstl, and Powers, 1991) of a differential area is equal to the sum of the radiance emitted and reflected from that area, plus the total radiance scattered and transmitted onto the area from other differential areas. An energy-balance equation is set up for each object in the medium. This gives Jf coupled differential equations, where Jf is the number of objects. Although the accuracy of this method is high, in order to use it one must know the detailed position, orientation, and shape of each object in addition to its scattering properties. This is seldom the case in either remote-sensing or laboratory applications, where a statistical description of the properties of the scatterers is usually all that is available. Hence, the method will not be considered further in this book. 7.FA. The doubling method This numerical method was developed by Van de Hulst and is widely used in calculations of radiative transfer in planetary atmospheres (e.g., Hansen and Travis, 1974). It is efficient in computer time and is especially useful when the particle phase function is anisotropic. The polarization as well as the intensity can be readily calculated. The concept is simple, although its practical realization may be complicated. Let a layer of particles have a reflectance R and transmittance T, where R is a matrix operator that describes the fraction of light incident from a certain direction d 0 that is reflected from the layer into various directions in the hemisphere containing the source, and T is the operator describing the fraction transmitted through the layer into directions in the hemisphere opposite the source.
175
176
7. Propagation in a nonuniform medium
The reflectance of two identical layers (Figure 7.5) may be written symbolically, R' = R+TRT + TRRRT+ • • • = R+TRT(1 + RR+
) = R+TRT(1 - R R ) ~ \
and the transmittance, T' = TT + TRRT + TRRRRT+ • • • = TT(1 + RR+
) = TT(1-RR)" 1 ,
where the superscript ( - 1 ) denotes the inverse operator. Once the operators T and R are specified, the inverse operator (1 —RR)" 1 and the reflectance and transmittance operators of the double layer can be found. The calculation starts with a layer that is so thin that its reflectance R and transmittance T can be written down by inspection. A typical optical thickness might be 2" 25 . A second, identical layer is added, and R' and T' of the doubled layer are calculated using the computer algorithm. This layer is again doubled; the R' and T' become the R and T of the new system, and the algorithm calculates the R' and T of the quadrupled layer. The process continues until the desired thickness is obtained. For instance, if the properties of a layer of optical thickness 2+25 arc desired, only 50 repetitions of the algorithm are required. The technique is readily modified to allow calculations of layers of differing properties. Figure 7.5. Schematic of the doubling method.
\
7.F. Solution of radiative-transfer problems
111
7.F.5. The multistream method
This is a technique for obtaining approximate solutions of the radiative transfer equation. The sphere of all propagation directions ft is broken up into Jf regions of solid angle Aft^, which need not be equal. The equation of radiative transfer, equation (7.21), is integrated in solid angle over each of the regions Aft. and each resulting equation divided by Aft., giving JK equations of the form cos
/ 'A
^ Air/ An / 4
(7.48)
Define the following average quantities for the /th zone:
* ATT f
=
cos^rfft,
A I T /An *X*>") *n •
(7.50)
( 7 - 52 )
Note that p^^ is the fraction of the radiance traveling in the direction of the center of region i that is scattered into region / . In (7.48), /(T,ft) is replaced by its average value lj<j\ which may then be taken out of the integrals over angle. Then the equation for
178
7. Propagation in a nonuniform medium
the intensity in the /th directional region becomes
^
^
+ ^r/(T),
(7.53)
where the region Cl^ contains the direction to the collimated source, and / takes all integer values between 1 and J^. Thus, the partial integrodiflferential equation (7.21) is replaced by a series of J^ linear, first-order, coupled differential equations that are amenable to numerical solution. In principle, the mesh can be made as fine as computer time allows, and the calculations can approximate the exact solution as closely as desired. If Jf is small enough, the equations can be solved analytically. In particular, if J^ = 2, the method is known as the two-stream or Schuster-Schwarzschild method (Schuster, 1905). It will be used extensively in the remainder of this book. The multistream method was used by Chandrasekhar (1960) in his classic treatise on radiative transfer. The directional sphere is divided into jy regions in polar angle #, where JV is an even integer. The boundaries are given by the zeros of P^(cos #), the Legendre polynomial of order JV. It was shown by Gauss that this choice of fixing the boundaries results in an approximation that is accurate to order 2JV, and the technique is known as the method of Gaussian quadratures. For example, for isotropic scatterers [pig) = 1] and S?T = 0, equation (7.21) has a solution of the form 7(T,£1) = ssfe~aT/[I + acos # ] + &e~T/tl0/[l + (l/iJL0)cos'&]. However, this solution cannot be made to satisfy the boundary conditions if the constants s/9 &, and a are the same for all directions. In Chandrasekhar's method the solution is forced to be of this form, but the "constants" are different within each Gaussian region. 7.F.6. The Eddington approximation In this method the solution of the radiative-transfer equation is assumed to be of the form /(r, H) = / 0 (T) + 71(r)cos #. This solution is substituted into (7.21), and the resulting equation is integrated over solid angle, giving one equation in / 0 (T) and / / T ) . Next this solution
7.F. Solution of radiative-transfer problems
179
is multiplied by cos d, substituted into (7.43), and integrated over 12, giving a second equation for / 0 (r) and / / T ) . Frequently ^ ( T ) is isotropic and independent of ft. Assuming that this is the case, the resulting equations are
(7.54a) and - \ ^ -
= -\h{T) + We-^,
(7.54b)
where £ is the cosine asymmetry factor. The two equations may then be solved simultaneously. 7.F.7. Successive approximations
In equation (7.21) the term on the left and the first term on the right-hand side of the equation can be put into the following form:
Hence, the radiative-transfer equation can be written
- T
Integrating both sides between r and oo gives T/COS#
Xe
where the boundary condition that /(r,ft) must remain finite as r -> oo has been used. To use the method of successive approximations, a trial solution for 7(T',ft') is inserted into the double integral on the right-hand side of (7.55), and a new solution is calculated, either analytically or numerically. This procedure may be repeated until the desired accuracy is obtained.
180
7. Propagation in a nonuniform medium 7.F.8. The method of embedded invariance
This technique allows one to obtain exact expressions for the reflectance and emittance of a semiinfinite medium in certain physically interesting cases without explicitly solving the radiative-transfer equation. It relies on the fact that the reflectance of, or thermal emission from, an infinitely thick layer of particles is not changed by adding a thin layer of identical particles. This will be discussed in further detail and used in Chapters 8 and 13.
8 The bidirectional reflectance of a semiinfinite medium
8.A. Introduction
In Chapters 8, 9, and 10, exact expressions for several different types of reflectances and related quantities frequently encountered in remote sensing and diffuse reflectance spectroscopy will be given. Next, approximate solutions to the radiative-transfer equation will be developed in order to obtain analytic evaluations of these quantities. As we discussed in Chapter 1, even though such analytic solutions are approximate, they are useful because there is little point in doing a detailed, exact calculation of the reflectance from a medium when the scattering properties of the particles that make up the medium are unknown and the absolute accuracy of the measurement is not high. In most of the cases encountered in remote sensing an approximate analytic solution is much more convenient and not necessarily less accurate than a numerical computer calculation. In keeping with this discussion, polarization will be ignored until Chapter 14. This neglect is justified because most of the applications of interest involve the interpretation of remote-sensing or laboratory measurements in which the polarization of the incident irradiance is usually small. Although certain particles, such as Rayleigh scatterers or perfect spheres, may polarize the light strongly at some angles, the particles encountered in most applications are large, rough, and irregular, and the polarization of the light scattered by them is relatively small (Chapter 6) (Liou and Scotland, 1971). Hence, to first 181
182
8. Bidirectional reflectance: semiinfinite medium
order, both the incident radiation and scattered radiation may be assumed to be unpolarized. 8.B. Reflectances
In this book we wish to develop formalisms that will allow the estimation of properties of a medium from the way it scatters or emits electromagnetic radiation from its upper surface. The terms reflectance and reflectivity both refer to the fraction of incident light scattered or reflected by a material. Although they are sometimes used interchangeably, reflectance has the connotation of the diffuse scattering of light into many directions by a geometrically complex medium, whereas reflectivity refers to the specular reflection of radiation by a smooth surface. Reflectivity was discussed in detail in Chapter 4, and reflectance will be the topic of the next several chapters. There are many kinds reflectance, depending on the geometry, so that the term must be appropriately qualified to be unambiguous. In modern usage (Nicodemus, 1970; Nicodemus et al., 1977) the word is preceded by two adjectives, the first describing the degree of collimation of the source, and the second that of the detector. The usual adjectives are directional, conical, or hemispherical. For example, the directional-hemispherical reflectance is the total fraction of light scattered into the upward hemisphere by a surface illuminated from above by a highly collimated source. If the two adjectives are identical, the prefix bi- is used. Thus, the bidirectional reflectance is the same as the directional-directional reflectance. The various reflectances and the symbols that will be used to represent them in this book are summarized in Table 8.1. In reality, all measured reflectances are biconical, because neither perfect collimation nor perfect diffuseness can be achieved in practice. However, many situations in nature approach the ideal sufficiently that the other quantities are useful approximations. To give several examples: Because the sun subtends only 0.5° as seen from earth, sunlight can be treated as collimated in most applications, whereas light scattered by clouds on an overcast day is nearly diffuse. Thus, on a clear, sunny day, the eye looking at the ground perceives bidirectional reflectance, whereas on an overcast day it perceives hemispherical-directional reflectance. A photosensitive device on a high-altitude
8.B. Reflectances
183
aircraft or spacecraft measures the bidirectional reflectance of the ground or clouds. The temperature of the surface of a planet is determined by a balance between the thermally emitted infrared radiation and the sunlight absorbed by the surface; the latter quantity is the difference between the incident sunlight and the product of the sunlight and directional-hemispherical reflectance. The reader is invited to think of other examples of different types of reflectance. Even within the general framework of definitions given earlier there is still a certain amount of arbitrariness in the way the various reflectances can be defined. For instance, a certain type of reflectance may be defined either in terms of power per unit surface area of the medium or in terms of radiances, which are power per unit area perpendicular to the direction to the source or detector. In general, I have tried to use definitions that are the most intuitively obvious or those that result in the simplest mathematical expressions for the reflectances. Nicodemus et al. (1977) have suggested a system of definitions and notation that is in wide use. Table 8.1. Types of reflectances and their symbols Bidirectional reflectance, r
> rdd
CtlMllg UJllllllclllUll Ul UCIHCUlUi
*
Directionalconical reflectance, rdc
Directionalhemispherical reflectance, hemispherical reflectance, hemispherical albedo, plane albedo,
Biconical reflectance,
Conicalhemispherical reflectance,
T
Decreasing collimation of source
Conicaldirectional reflectance, r
cd
Hemisphericaldirectional reflectance, r
r
cc
Hemisphericalconical reflectance,
hd
r
hc
Bihemispherical reflectance, spherical reflectance, spherical albedo, Bond albedo, r
Indeterminate: Diffusive reflectance, r0 Kubelka-Munk reflectance, r'o
h> rdh
s> rhh
184
8. Bidirectional reflectance: semiinfinite medium
Although I have tried to follow that system where convenient, some of the definitions used here differ from those of Nicodemus and associates. Sometimes a symbol proposed by those authors is already being widely used to represent a different quantity. For instance, Nicodemus and associates propose using p as the symbol for reflectance. However, p is already commonly used to denote mass density; thus, I use r for reflectance. In general, two subscripts are added to r to indicate the degrees of collimation of the source and detector (e.g., rdh = directionalhemispherical reflectance). However, although this terminology is precise, it is unwieldy. Hence, whenever the meaning is unambiguous, the following contractions will be used: (1) The bidirectional reflectance, denoted by rdd, occurs so often in this book that the simple symbol r with no subscript will be used for it most of the time. (2) Most commercial reflectance spectrometers measure the directionalhemispherical reflectance, denoted by rdh, but this quantity is also widely used in astrophysics, where it is called the hemispherical albedo and the plane albedo. Thus, for convenience, it will frequently be referred to simply as the hemispherical reflectance and denoted by rh. (3) In Chapter 10 it will be shown that the bihemispherical reflectance, denoted by rhh, is equivalent to a quantity known in astrophysics as the spherical reflectance, spherical albedo, and Bond albedo. Hence, the bihemispherical reflectance will be called the spherical reflectance and denoted by rs.
An approximate analytic expression for the bidirectional reflectance of a particulate medium of infinite thickness is derived and discussed in detail in this chapter. Further applications are discussed in Chapter 9. The reflectances that involve integration over hemispheres are discussed in Chapter 10. In the remote sensing of bodies of the solar system, several additional types of reflectances are in use and are referred to as albedos and phase functions. These will be defined and considered in detail in Chapter 10 also. 8.C. Geometry and notation
The geometry and nomenclature that will be used in the remainder of the book are defined in this section. The geometry is illustrated schematically in Figure 8.1. Collimated light (irradiance) / from a source of radiation is incident on the upper surface of a scattering
8.C. Geometry and notation
185
medium. The normal to the surface N is parallel to the z axis, and the incident light makes an angle I with N. The light interacts with the medium, and some of the rays emerge from an element A^4 of the surface traveling toward a detector in a direction that makes an angle e with N. The plane containing the incident ray and N is the plane of incidence, and that containing the emerging ray and N is the plane of emergence. The azimuthal angle between the planes of incidence and emergence is \fj. The angle between the directions to the source and detector as seen from the surface is the phase angle g. The plane containing the incident and emergent rays is the scattering plane. If the planes of emergence and incidence coincide (I/J = 0 or 180°), their common plane is called the principal plane. A commonly used notation, which will be followed in this book, is to let )it = cose, JJLO = cos i.
(8.1) (8.2)
In general, three angles are needed to specify the geometry. The angles usually used in terrestrial remote sensing or laboratory applications are i, e, and if/. However, most planetary applications specify i, e, and g, the reason being that in a spacecraft or telescopic image of a planet, g is very nearly constant over the entire image. Occasionally, the scattering angle 0 may be used instead of g. Figure 8.1. Schematic diagram of bidirectional reflectance from a surface element A^4, showing the various angles. The plane containing J and I is the scattering plane. If the scattering plane also contains N, it is called the principal plane.
186
8. Bidirectional reflectance: semiinfinite medium
When A A is located on the surface of a spherical body, the geometry may be specified by luminance coordinates, which consist of the luminance longitude A, luminance latitude L, and phase angle g. This spherical coordinate system, which has nothing to do with the geographic coordinates on the planet, is illustrated in Figure 8.2. The luminance equator is the great circle on the surface of the planet containing the sub-source point S and sub-observer point O. The luminance axis is the diameter perpendicular to the luminance equator. The phase is positive if S is to the left of O, and negative if to the right, as seen by the observer. The prime luminance meridian passes through O. Luminance east and positive longitudes are to the right of O; luminance west and negative longitudes are to the left of O. The relationships among the various coordinate systems may be found using the law of cosines for spherical triangles. Suppose a triangle whose sides are great circles on the surface of a sphere has sides A, B, and C, which are separated by interior angles a, b, and c (Figure 8.3). The law of cosines states that cosC = cosAcosB + sinAsinBcosc.
(8.3)
Identical relations hold between the other sides and angles. Figure 8.2. Schematic diagram of luminance coordinates on a spherical planet. N
8.D. Radiance of a semiinfinite medium Applying the law of cosines to the triangle SO A A in Figure 8.2 gives cos g = cos I cos e + sin I sin e cos i//. (8.4) Applying the law to the triangle SEA A gives cos I = cos(A + g)cos L, (8.5) and to the triangle OEAA, (8.6) cos e = cos A cos L. When g > 0, the terminator is located along the meridian of luminance longitude where I = rr/2, corresponding to A = TT/2— g, and the bright limb occurs at e = A = — IT / 2 . 8.D. The radiance at a detector viewing a semiinfinite medium In most laboratory and remote-sensing applications the quantity of interest is the radiance received by a detector viewing a horizontally stratified, optically thick medium of particles that may scatter, absorb, and emit. The relation between the radiance at the detector and the radiance field in the medium may be calculated from the equation of radiative transfer. The geometry is indicated schematically in Figure 8.4. Let z be the vertical distance on the axis perpendicular to the planes of stratification. The distribution of particles with altitude z is arbitrary, except that the particle density N -> 0 as z -> <», corresponding to r -> 0. The particles in the space below the T = 0 level are characterized by the radiative-transfer parameters E(z), S(z), K(z), G(z,fl',ft), and F(z,ft), as defined in Section 7.4. The space above the r = 0 level is empty, except for a distant source of collimated irradiance / that illuminates the medium and a detector that views the medium. The Figure 8.3. Spherical triangle.
187
188
8. Bidirectional reflectance: semiinfinite medium
sensitive area of the detector is A0, and it responds to light that is incident only within a small solid angle Ao>. The light incident on the detector from the medium is interpreted as if it were emitted from an area A A formed by the intersection of Ao> with some surface, which is interpreted by a distant observer as the apparent surface of the medium. However, the radiance at the detector actually comes from light scattered or emitted by all the particles in the medium within the detector field of view Aw. If the distribution of scatterers with altitude is a step function, the apparent surface is the actual upper surface of the medium. If the altitude distribution is nonuniform, the apparent surface is usually defined to be at the level corresponding to T = 1. Consider a volume element d*V = &2Aco d$l located within Aco at an altitude z in the medium and a distance 3£ from the detector. This volume element is bathed in radiance J(z,ft')dft' traveling within solid angle dQI about direction ft'. Thus, an amount of power is scattered by the particles in )f47rG(z,n\n)I(z,af)dnf into unit solid angle about the direction ft between d'V and the detector. In addition, an amount of power F(z,Cl)d^ is emitted from dJT per unit solid angle toward the detector. Figure 8.4. Geometry of scattering from within a particulate medium. The nominal surface of the medium is the x-y plane.
8.D. Radiance of a semiinfinite medium
189
Now, the solid angle of the detector as seen from d'V is La/312. The radiance scattered and emitted from d"V toward the detector is attenuated by extinction by the particles between d"V and the detector by a factor e~r/lx before emerging from the medium. Hence, the power from d"V reaching the detector is 4
S(Z)
E{Z)
+ ^(r,n)]e-^^,
(8.7)
where we have put dz = \x d$l = — dr/E. The total power reaching the detector is the integral of dPD over all volume elements within Aw between z = — <*> and +oo, or, equivalently, between r = oo and 0. The radiance ID at the detector is the power per unit area per unit solid angle. Thus, 1 W(T)
r
^
Kir
,
r\ '
r
/
\1 J
-T/
dr ^
(8.8) But from the radiative-transfer equation (7.43), the factor in brackets is
Because /(T,ft) must remain finite as r
Thus, the radiance at the detector is equal to the radiance emerging from the medium at the r = 0 level toward the direction of the
190
& Bidirectional reflectance: semiinfinite medium
detector, provided that the detector always views an area A A on the apparent surface that is much smaller than either the size of the medium or the illuminated area of the apparent surface. Hence, the problem of finding the radiance at the detector from an optically thick medium is equivalent to determining the radiance leaving the medium in the direction of the detector. Now, ID = 7(0, fl) is the power emitted by the medium into unit solid angle about a direction making an angle e with the vertical per unit area perpendicular to this direction. The projection of this unit area onto the apparent surface has area sec t. Hence, the power emitted into unit solid angle from unit area of the apparent surface of the medium is \p Y
8.E. Empirical expressions: the Lambert and Minnaert reflectance functions 8.E.L Lambert's law (the diffuse-reflectance function) Before deriving any reflectances, two empirical expressions that are widely used because of their mathematical simplicity will be introduced. The first of these is known as Lambert's law and is the simplest and one of the most useful analytic bidirectional-reflectance functions. Although no natural surface obeys Lambert's law exactly, many surfaces, especially those covered by bright, flat paints, approximate the scattering behavior described by this expression. Lambert's law is based on the empirical observation that the brightnesses of many surfaces are nearly independent of e and I/J and on the fact that the brightness of any surface must be proportional to cos I. Because light scattering is a linear process, the scattered radiance I(l,e,il/) = Jr(l,e,i/j) must be proportional to the incident irradiance per unit surface area J/L0. The bidirectional reflectance of a surface that obeys Lambert's law is defined to be rL(l, e,i//) = KLcos I, where KL is a constant.
(8.11)
8.E. Lambert and Minnaert reflectance functions
191
The total power scattered by unit area of a Lambert surface into all directions of the upper hemisphere is
TT/2
/
6=
rlir
I
JKL cos I cos e sin e d e di/j =
0 «V = 0
TTJKLIJL0.
Because the incident power per unit area is Jfi0, the fraction of the incident irradiance scattered by unit area of the surface back into the upper hemisphere is the Lambert albedo AL = PL/J/JL0 = KLTT, SO
that KL =
AL/v.
Thus, Lambert's law is I(i,e,tlj)
= ^ALfji0,
(8.12)
and the Lambert reflectance is ^ L ( ^ ^ ^ ) = -^LMO-
(8.13)
Note that AL is the directional-hemispherical reflectance of a Lambert surface. A surface whose scattering properties can be described by Lambert's law is called a diffuse surface or Lambert surface. If Ah = 1, the surface is a perfectly diffuse surface. Lambert's law is widely used as a mathematically convenient expression for the bidirectional reflectance when modeling diffuse scattering. In fact, it does provide a reasonably good description of the reflectances of high-albedo surfaces, like snow or flat, light-colored paint, but the approximation is poor for dark surfaces, such as soils or vegetation. Figure 8.15 compares equation (8.13) with (8.89), which accurately describes scattering from a variety of particulate surfaces, for the cases of high and low albedos. 8.E.2. Minnaert's law The power emitted per unit surface area per unit solid angle by a medium obeying Lambert's law is Y= Jrh/jL = (J/ir)AhiJLQiJL. In this expression the reflected power per unit area is proportional to (/XQ/X)1. Minnaert (1941) suggested generalizing Lambert's law so that the power emitted per unit solid angle per unit area of the surface be
192
8. Bidirectional reflectance: semiinfinite medium
proportional to (fjiojjL)v. This leads to a bidirectional reflectance function of the form rM(l,e,i!,) = AMnW-\ (8.14) where AM and v are empirical constants. Equation (8.83) is known as Minnaert's law, v is the Minnaert index, and AM is the Minnaert albedo. If v = 1, Minnaert's law reduces to Lambert's law, and AM = AL/7T.
It is found empirically that Minnaert's law approximately describes the variation of brightness of many surfaces over a limited range of angles (cf. Thorpe, 1973; Veverka et al, 1978c; McEwan, 1988). However, a number of difficulties attend the use of Minnaert's law. First, the general law breaks down completely at the limb of a planet where e = 90°:If v < 1 the calculated brightness becomes infinite, and if v > 1 the limb brightness is zero, neither of which agrees with observations of any real body of the solar system. Second, the Minnaert parameters are not related in any obvious way to the physical properties of the surface. Third, the Minnaert "constants" AM and v both depend on phase and azimuth angle (Hapke, 1981). 8.F. The diffusive reflectance 8.F.L Equations
To introduce the two-stream method of solution of the equation of radiative transfer, we will now derive one of the simplest and most useful expressions in reflectance theory, the diffusive reflectance. In spite of the fact that this expression is an approximation, its mathematical simplicity allows analytic expressions to be obtained for a variety of scattering problems. These expressions provide surprisingly good first-order estimates in many applications. In a very large number of papers in the literature involving the scattering of light within and from planetary surfaces and atmospheres, high-precision numerical techniques have been used to obtain answers that could have been calculated to a satisfactory degree of accuracy much more conveniently using the diffusive reflectance. In addition, it will be seen that the diffusive reflectance appears in the mathematical expressions for several other types of reflectances, or that other reflectances reduce to it at special angles. Thus, the diffusive reflectance is representative of solutions of the radiative-
8.F. The diffusive reflectance
193
transfer equation. An expression known as the Kubelka-Munk equation, which is widely used to interpret reflectance data, will be seen in Chapter 11 to be a form of the diffusive reflectance. The problem to be solved is the following: Radiance is incident on a plane corresponding to r = 0 that separates an empty upper half-space from an infinitely thick lower half-space filled with particles that scatter and absorb light. The collimation of the incident light is unspecified, but in most applications the radiance is assumed to be uniformly distributed over the entire upper hemisphere. The simplifying assumption that characterizes the diffusive-reflectance formalism is that the incident radiance does not penetrate into the medium, but acts as if the upper boundary were covered by an invisible membrane that selectively scatters the incident light uniformly into all downward directions. Because it is assumed that there are no collimated or thermal sources of radiance within the medium, the source function in the radiative-transfer equation <^"(r,ft) = 0. The problem is to find the radiance inside the medium and the scattered radiance emerging from the r = 0 plane in the upward direction into the upper half-space. With these conditions the multistream approximation for the / t h region, equation (7.53), becomes
We will use the two-stream approximation, so that JV = 2. The two regions in the directional solid angle are chosen to be the upward-going hemisphere, denoted by / = 1, and the downward-going hemisphere, denoted by / = 2. Then AQ1 = A£12 = 2TT, and equations (7.49) and (7.50) become Lti = 7s— / 2ir JO
a, = ~— f
cos # 2TT sin fidfi = -^, 2
cos # 2TT sin &dft = - ~-.
The angular-scattering properties of the particles will be characterized by the hemispherical asymmetry factor /3, defined in Chapter 6, so that
194
& Bidirectional reflectance: semiinfinite medium
Hence, equations (8.15) become
The particle scattering parameters w and /? are assumed to be independent of T. Let the downward radiance emerging from the "membrane" at r = 0 be / 0 . Then the boundary conditions on these equations are that IX(T) and / 2 ( T ) remain finite as r ->o° and that at the surface the downward flux be /2(0) = / 0 . 8.F.2. Isotropic scatterers If the particles of the medium scatter light isotropically, then /3 = 0, and equations (8.16) are
-\l£ i%
=
-h + ^(h + h),
= -I2 + %(l2 + h),
(8.17a) (8.17b)
The solution is facilitated by letting
(8.18a)
A
(8.18b)
and
so that I^cp
+ Acp
(8.19a)
/ 2 =
(8.19b)
and Physically, cp(r) is the directionally averaged radiance in the medium. By alternately adding and subtracting (8.17a) and (8.17b), the twostream equations can be put into the form w)(p,
(8.20a) (8.20b)
8.F. The diffusive reflectance
195
Differentiating (8.20b) and substituting the result into (8.20a) gives
0
= 4(1-*)?.
(8.21)
It may be readily verified by direct substitution that the general solution of (8.21) is cp(r) = srfe-2yT + &e2yr, (8.22a) where r
= (l_H;)1/2
(8.22b)
is the albedo factor, and sf and & are constants to be determined by the boundary conditions. Then, from (8.20b), I ^7
yr
-
(8.22c)
Converting back to Ix and / 2 , I^T) = cp + A
(8.23a)
hi?) = 9 ~ A 9 = -^(1 + r ) ^ " 2 r r + ^ ( 1 " y)^ 2yT , (8.23b) In an infinitely thick medium,
/ 2 (r) = / 0 e - 2 ^ . (8.24b) Now, the total downwelling power emerging from unit area of the membrane" at the surface of the medium is Pm = r
/
Iocos# 2*77sin$dd = TT/0,
which is the same as the total power per unit area incident on the upper surface of the medium. Similarly, the total scattered power emerging in all upward directions from unit area of the surface is Pem = f 77/2 / 1 (0)cos # 2 77 sin ddft = TTI^O) = TTIOT^-
.
The diffusive reflectance, which will be denoted by r 0, is the ratio
196
8. Bidirectional reflectance: semiinfinite medium
\n. Thus, (8.25) Let us explore some of the properties of r 0. The diffusive reflectance ro(w) is plotted against w in Figure 8.5. Note that ro(O) = 0 and r o (l) = l. Using Taylor's theorem, equation (8.25) may be expressed as a power series in w9
ro(M0-4* + ^
+
^+....
(8.26)
In this series the term proportional to w' is the contribution of the /th order of multiple scattering to r 0. Thus, the contribution of single scattering is H>/4, and that of doubly scattered light is w 2 /8, and so on. The total contribution of multiple scattering is ro-w/4 = r o [ l - ( l + r o )~ 2 ]. For surfaces of very high albedo, multiple scattering is seen to contribute about 75% of the scattered light. For small values of w, only single scattering is important, and the curve is the straight line r o (w)-H>/4. As w increases, the contribution of higher-order scattering becomes significant, and r0 increases nonlinearly, until at the point r o (l) = 1 the slope is infinite. Equation (8.25) is readily solved for y and w in terms of r 0: -r 0 ), (8.27) 2
.
(8.28)
Figure 8.5. Diffusive reflectance as a function of single-scattering albedo.
0
0.2
0.4
0.6
0.8
1
&F. The diffusive reflectance
197
These expressions are useful for rapid estimates of w or y from measured values of reflectance. The answers to a number of interesting questions may be estimated using the diffusive-reflectance model. For example, how many scatterings, on the average, does a photon undergo before being scattered out of a medium? This quantity can be calculated by noting that the total fraction of light absorbed by the medium is 1 — r 0 and that the fraction absorbed at each scattering is 1 - w. Therefore, the average number of scatterings, Jt, is
Thus, in a powder with particles having w = 0.5, r0 = 0.18, and the average photon emerging from the surface has undergone ^ # = 1 . 7 scatterings. If w = 0.9, r0 = 0.5 and at« 5; if w = 0.99, r 0 = 0.8 and
8.F.3. The Lambert-diffusive-scattering law The diffusive approximation may be combined with Lambert's law by setting the total incident power per unit area TT/0 = / / i 0 and assuming that the emergent radiance is independent of e. Then the Lambertdiffusive expressions for the bidirectional and hemispherical reflectances are, respectively, 1
and ~" r 0 •
If the surface is Lambertian, but the incident radiance is diffuse and the same in all directions, then the Lambert-diffusive spherical reflectance is
8.F.4. Anisotropic scatterers: the similarity relations If j8 # 0, then equations (8.16) may readily be solved using the same procedure as for isotropic scatterers. However, it is instructive to ask if it is possible to transform equations (8.16) into the same form as (8.17). That is, we seek quantities T* and w* such that (8.16) can be
198
8. Bidirectional reflectance: semiinflnite medium
written
-\§^
= -h^(h =
\w*
+ h),
(8.29a)
-h + ^:(h + h)-
(8.29b)
Equating the coefficients of It and I2 in (8.29) and (8.16) gives the following simultaneous equations; H>*T* = ( 1 - J 3 ) H > T ,
T* - 2WT
= T
" 2(* + i 8 ) " ^ .
Solving these yields T* = ( 1 - / 3 H > ) T ,
(8.30a)
"*-T=fiWw-
( 8 - 30b )
Equations (8.30) are known as similarity relations. Sometimes the cosine asymmetry factor £ is used instead of /3. Because equations (8.29) are of the same form as (8.17), solutions for media of anisotropic scatterers are the same as for isotropic scatterers, except that the quantities T and w are replaced by r* and H>*, respectively. Thus, (8.31a) (8.31b) where, for a semiinfinite medium, & = 0, sf = 2/ 0 / ( I + y*), and y* = [ l - w * ] 1 / 2 = [ ( l - w ) / ( l - i 8 w ) ] 1 / 2 ,
(8.32a)
so that y*T*
= [(i - w ) ( i _ 0H,)] 1 / 2 T .
(8.32b)
The diffusive reflectance is
^o(^) = ^ J .
(8.33)
When the particles are fully backscattering, p = - 1 and r 0 = (Vl +w> - V l - w)/(\/l + w + y/l — w). As in the case in which j8 = 0, /• 0(0) = 0 and r o (l) = 1, but when H>
8.G. The bidirectional reflectance
199
j8 = 1, then all of the radiance is scattered deeper into the medium and never escapes if any absorption is present. This, of course, is not a situation that is realizable in practice, and it shows that this formalism breaks down if |j8| is too close to 1. In general, when /3 > 0, r0 is smaller than its value for /3 = 0, and larger when /? < 0.
8.G. The bidirectional reflectance 8.G.I. Introduction The bidirectional reflectance of a medium is defined as the ratio of the scattered radiance at the detector to the incident irradiance. From the result of Section 8.D, this is equivalent to the radiance emerging in a given direction from a surface illuminated by a collimated irradiance per unit incident irradiance. The bidirectional reflectance of a particulate medium of infinite optical thickness will be derived in this chapter. Layered media will be considered in Chapter 9. We will begin by considering the simple case in which the singlescattering albedos of the particles are so small that multiply scattered light can be neglected. This will give a well-known expression, the Lommel-Seeliger law. Following that derivation, more general relations that include the effects of multiple scattering will be discussed in detail, and useful approximations given.
8.G.2. Single scattering: the Lommel-Seeliger law The contribution to the bidirectional reflectance of a semiinfinite, particulate medium by light that has been scattered only once can be calculated exactly from equation (8.8). By assumption, there are no thermal sources. Referring to equation (8.8) and to Figure 8.4, Because we the source function is ^ ( r , f t ) = Je~T/fJ/°w(T)p(T,g). are ignoring multiple scattering, then in equation (8.8) the integral / 47rJ p(T,ft / ,ft)/(r,fl')rfn = O. The total radiance IDs reaching the detector due to single scattering is then r.
(8.34)
If w and p are independent of z or r, as is often the case, then the
200
& Bidirectional reflectance: semiinfinite medium
evaluation of this integral is trivial, and gives (8.35a) When p(g) = 1, that is, when the scatterers are isotropic, equation (8.35a) is the LommelSeeliger law. This law has been generalized in (8.35a) to include nonisotropic scatterers. Except close to zero phase, this expression is a good description of the light scattered by lowalbedo bodies of the solar system, such as the moon and Mercury (Hapke, 1963, 1971), for which only light that has been scattered once contributes significantly to the brightness. Transforming to luminance coordinates [equations (8.5) and (8.6)], (8.35a) becomes cos(A
w
(8.35b)
4
Note that this expression is independent of luminance latitude L. To a fair approximation, the brightness of the moon at small phase angles is, in fact, independent of latitude (Minnaert, 1961; Hapke, 1971). The Lommel-Seeliger function cos(A + g)/[cos(A + g) + cos A] is plotted as a function of longitude for three phase angles in Figure 8.6. The surge in brightness near the limb is not observed on the moon because of the roughness of the lunar surface; effects of macroscopic surface roughness are discussed in Chapter 12. Figure 8.6. Lommel-Seeliger law versus luminance longitude for several values of the phase angle.
0.8
0.6
-
0.4
"
0.2
-90
-60
-30
0
Longitude (degrees)
30
60
90
&G. The bidirectional reflectance
201
8.G.3. The bidirectional reflectance of a medium ofisotropic scatterers 8.G.3.a. The two-stream solution with collimated source
In this section we will show how multiple scattering may be included in the calculation of the bidirectional reflectance. In order to illustrate both the power and the limitations of the two-stream method, this technique will be used to obtain an approximate solution to the radiative-transfer equation for a semiinfinite medium of isotropic scatterers. The exact expression for the bidirectional reflectance of such a medium will be found in the next section using the method of embedded invariance, and the two solutions will be compared. Lumme and Bowell (1981) have also derived an expression for the bidirectional reflectance; their approach is somewhat similar to that described in this book, although differing in several major details. The geometry is the same as in Figure 8.1. Collimated irradiance / is incident on a particulate medium. The space above the plane corresponding to r = 0 is empty, except for the source and detector, and the volume below this plane contains particles that both scatter and absorb. Thermal emission is assumed to be negligible. The properties of the medium are described by the nomenclature of Chapter 7, and w and pig) are assumed to be independent of r. The radiative-transfer equation in the form of (7.21) will be used, in which pig) = 1, y r = 0, and <^(T, g) = Jwe~r/tlQ. A combination of the two-stream method and the method of successive approximations will be used. First, an approximate expression for /(r,ft) will be obtained using the two-stream method. Next, this first-order solution will be substituted back into (7.56) to give a second-order solution. Finally, the reflectance will be calculated by inserting the second-order solution into (8.9). As in Section 8.F, the two-stream form of the equation of radiative transfer is obtained by putting JT = 2 in equation (7.52). The upwardgoing hemisphere is denoted by subscript / = 1, and the downwardgoing hemisphere by / = 2. Then Aftj = Aft2 — 2?r, \xx = \, fx2 = — \, pi^ = 1, and equation (7.53) becomes the set of two equations
(8-36b) The boundary conditions are that the radiance must remain finite IW
= =
-/2
+
T(/I
+ /
2)
+ /
^"
T / / 1
°-
202 & Bidirectional reflectance: semiinfinite medium everywhere and that there be no sources of diffuse radiation above the upper surface, so that at r = 0 the downward-going diffuse radiance J2(0) = 0. As in Section 8.F, put
2 dr 1 d
A~re
'
(^o.j/aj
(8.37b)
where y = Vl— w is the albedo factor. Differentiating (8.37b) and inserting into (8.37a) gives (8.38)
This equation has the solution cp(r) = s/e~2yr + ^e2yr + 9e~r/lL\ (8.39) where s/, &, and 9 are constants to be determined from the boundary conditions. Then, from (8.37b), A(p =
\ iff = " y ^ e ' 2 y T + y ^ e 2 y r - f~e'T/lx^
(8-40)
Because Ix and I2 must remain finite as T -> <*>, ^ = 0. Substituting (8.39) with 38 = 0 into (8.38) gives Mo " 2yr
2yr
+ ^ " T / / 4 ° ) + J^-e'T/iL\
(8.41)
r/
Now, e~ and e~ ^° are independent functions of r. Hence, the only way (8.41) can be true for all values of T is if the coefficients of these functions are separately equal. Equating the coefficients of 6 -T/MO
gives
Equating the coefficients of e'2yT gives an identity, which confirms that (8.39) is the solution of (8.38). Converting back to Ix and I2 gives 1 (8.43a) 11+
(8.43b)
8.G. The bidirectional reflectance
203
From equation (8.37b), the boundary condition that / 2 (0) = 0 is equivalent to
i
^
(8.44)
Using either form of the boundary condition at T = 0 gives
(l-y)2v0y
(845)
Because
= w l w
T/COS#
I
/
^ \
l/cos#+2y c 7
1/cos*-
+
Substituting for & and s/ from (8.42) and (8.45), we obtain Jw
1
[
( 1 - 7 ) 2 ^ ( 1 + 2 ^ ) _ 2yr
\lL\-l)e~r/lL*\.
(8.46)
Equation (8.46) is the two-stream solution for the radiance within the medium. Note that it contains two terms. The second term on the right is proportional to the source term and is important only within a distance from the surface of a few times the extinction length 1/E, which is of the order of the particle separation. The first term depends on y/E, which can be much longer than 1/E if the particles have high albedos. From (8.9), the radiance at the detector is Id = 7(0, /JL), SO that the bidirectional reflectance is From (8.46), after a little algebra, the expression for the reflectance
204
& Bidirectional reflectance: semiinflnite medium Figure 8.7. Schematic diagram of the five first-order changes in the scattered radiance caused by adding a thin layer of optical thickness AT to the top of an infinitely thick medium of bidirectional reflectance r.
(a)
(b)
8.G. The bidirectional reflectance
205
can be put into the form
8.G.3.b. The method of embedded invariance
Equation (8.47) is an approximate solution for the bidirectional reflectance of a semiinfinite medium of isotropic scatterers. The exact, Figure 8.7. (cont.)
206
& Bidirectional reflectance: semiinfinite medium
rigorous expression may be found without solving the equation of radiative transfer using a principle called embedded invariance. This method, which was developed by Ambartsumian (1958), is based on the fact that the reflectance of a semiinfinite medium does not change if a thin additional layer is added to the top of the original surface. Ambartsumian's solution will be reproduced here and compared with the two-stream solution. Let the bidirectional reflectance of the medium be r(i, e,i//). Then, because each particle scatters the same fraction of incident light equally into all directions of azimuth i//, r(L, e,i/r) must be independent of iff. Hence, t/r may be ignored as an independent variable, and the reflectance denoted by r(i, e), or equivalently by r(ju0, M). Let a thin layer of thickness Az of identical particles be added to the top of the layer. Assume that its optical thickness AT = - £Az is so small that interactions of light with the layer involving powers of AT greater than 1 can be ignored. Then the layer will cause five distinct changes that are each proportional to AT in the scattered light. These changes are shown schematically in Figure 8.7: (1) Light passing through the added layer is reduced by extinction (Figure 8.7a), once by a factor exp( — AT//X 0 ) on the way in, and once by a factor exp( — AT//X) on the way out. Thus, the emergent radiance Figure 8.7. (cont.)
8.G. The bidirectional reflectance
207
due to this effect alone is
so that the first-order change in the emergent radiance is ''6)-
(8 48a)
-
(2) The added layer scatters an additional amount of light toward the detector (Figure 8.7b). Consider a cylindrical volume coaxial with a scattered ray emerging from the layer in the direction toward the detector. The volume has cross-sectional area aQs and length Az//x. Incident light / will be scattered by any particle with its center in this volume, and this scattered radiance will be added to the emergent radiance scattered by the lower medium. The increase in radiance due to this effect is
A/2 = / M ^ ^ = / £ ^ .
(8.48b)
(3) Light scattered by the added layer in the downward direction is an additional source of illumination of the lower medium (Figure 8.7c). The radiance scattered into a given downward direction such that it is incident on the medium from a direction making an angle I1 with the normal is J(W/4TT)(AT/IJ!0), where IJ!0 = COS Lf. The light within a small increment of solid angle dft'. = sin i dt1d\\t\ = — d/j!0 dty\ is dl[ = J(w/4ir)(Ar/ij!0)dClfi, where ty[ is the angle between the azimuths^of the scattered ray and the exit ray. An amount dl-r(i\ e) is scattered by the lower medium into a direction toward the detector with emergent angle e. To find the total additional light reflected, this quantity must be integrated over ft'.. Hence,
M3=f = 4 AT/1
rr{i\e,)J^-^r(Mf0>M)^.
sin (8.48c)
(4) Light scattered upward by the lower medium illuminates the added layer (Figure 8.7d). The light scattered by the medium into an increment of solid angle dfte> = sin ede'dfo around a direction having emergent angle e is dle> = /r(i, e')dfte>. This light illuminates a
208
8. Bidirectional reflectance: semiinfinite medium
cylindrical volume akzQs/fji in the layer coaxial with a direction having emergent angle e. The light scattered within the volume into Integrating this over d£lj gives this direction is dlc>(w/4TT)(AT/fi).
4^irdI>f=z
J
w
Jr(<,,e')j-
AT
—sine'de'd^,. (8.48d)
where /x' = cos e'. (5) Light scattered upward by the lower medium is scattered back down by the added layer and provides an additional source of illumination of the lower medium (Figure 8.7e). An amount of light dl^ = is scattered by the lower medium into solid angle ie"dilje» about a direction having polar angle e" and azimuth «/r e». This light illuminates a cylindrical volume in the added layer containing particles that scatter light downward toward the lower medium and illuminate it from a direction having an angle of incidence i" and azimuth ^ with light of intensity dlu = dlt»(w/4TT)(AT/(i" 0) dCly, where /x,"0 = cos I"', and dCl^ = sini"di"dtyf. An amount of light dlur{l'\e) is scattered into a direction having emergent angle e. The total amount of light due to this effect is
A/ 5 =r
f
r(i",e)dlu
X sin e" d e" dtyj sin I" d l" The integrals over the four variables can be separated and integrated independently. Carrying out the integrations over azimuth, and factoring the integrals over 1 and i'\ gives / Vo=o
L // = o
(8.48e)
&G. The bidirectional reflectance
209
The sum of the changes A/ x through A/ 5 given by equations (8.48) must be zero. Hence,
w
w ^0
z
f/1 r(^, In the last term, JK,"0 and jtt" have now become simply dummy variables of integration and may be replaced by /A'O and /1,', respectively. Multiplying through by ATT/JL/W gives
(8.49)
Let Mo
Then (8.49) becomes
Written in this form it is seen that the function & is symmetric with respect to /A0 and /JL. Furthermore, the first term in brackets is a function only of /x, and the second term is the identical function of /JL0 only. Denote this function by H(x), where x may represent either /x or /i 0 . Then & may be written ^(JU,0,JU,) = H(iio)H(fL), where
210
& Bidirectional reflectance: semiinfinite medium
is the solution of the integral equation
H(x) = l+^-xH(
(8.50)
o
where x' is a dummy variable of integration. Using this result in (8.49) gives (8.51) where H(x) is a function that satisfies (8.50), and x is either JJL0 or fi. Comparing (8.51) with (8.47), we see that the two solutions are of identical form, except that in (8.47), H(x) is approximated by (l + 2x)/(l + 2yjc). Note that (8.51) is an exact, general solution for r and makes no assumptions about the medium, other than that it is composed of particles that scatter light isotropically and independently of each other. 8.G.3.C. Properties, analytic approximations, and moments of the H functions Equation (8.51) is the exact solution for the bidirectional reflectance of a semiinfinite medium of isotropic scatterers. However, the solution is in terms of the nonlinear integral equation (8.50). Values of the H Figure 8.8. Function H(x) versus x for several values of w. Solid lines, exact solution; dashed lines, approximation [equation (8.55)]. The more exact approximation, equation (8.57), is indistinguishable from the exact solution at the resolution of this figure. 1
•
1
.
1
.
1
•
1
2.5 w=1 .00
;
2
^ "
"
0.9 H(x)
,
1.5
-g1
" :
^
^
^
^
^
^
05
^
•
1
Mi —
•'
•
m
m »•
0
0.5
Exant
" ;
Approximation i
.
i
0.2
.
i
.
i
.
i
0.4
0.6 x
0.8
' •*
-
.
8.G. The bidirectional reflectance
211
functions for isotropic scatterers have been tabulated in several places (cf. Chandrasekhar, 1960). They are plotted for several values of w in Figure 8.8. In this section, some of the properties of the H functions will be described and some useful analytic approximations derived. Note that H(x) = 1 for all values of x when w = 0. In that case, (8.51) reduces to the Lommel-Seeliger law, equation (8.34) with p(g) = l, which describes single isotropic scattering. When x -> 0, H(x) -> 1, showing that at glancing angles of incidence and emergence only single scattering is important in the reflectance, no matter what the value of w. For w > 0, the function has a logarithmically infinite slope at x = 0, but as x increases, the curve rapidly flattens and becomes almost a straight line whose slope increases monotonically with w. The /th moment of the H function is defined by
f(x)x>dx.
(8.52)
o The Oth moment, which is simply the integral of the H function, and also its average value over the interval between 0 and 1, may be found from equation (8.50) as follows:
Rearranging gives This quadratic equation has the roots Ho = 2(1 ± \/l - w )/w. The minus sign must be chosen, because HQ is finite as w -* 0. Thus, putting y = ]/l — w gives
#o = T^7-
(8-53)
212
& Bidirectional reflectance: semiinfinite medium
Alternatively, Ho may be expressed in terms of the diffusive reflectance r 0 as (8.54) J f o « l + r o. The first moment H1 has been calculated by numerical integration by Chamberlain and Smith (1970). We will now describe a few useful analytic approximations to the H functions. The first is the two-stream approximation from equation (8.47),
This approximation is plotted in Figure 8.8. It differs by less than 4% from the exact values everywhere, and it is better than that in most places. The observation that H(x) is almost linear over most of its range suggests that for certain purposes it may be approximated by a linear function of the form H(x) =* $/ + ^JC, in which the integral over x is required to equalHo exactly. This gives the condition sf + £8/2 = Ho. Because H(0) = 1, a possible choice might be to take J / = 1, giving i / ( j t ) ^ l + frojt. Because the slope of H(x) is infinite at JC = O, a better choice is to set the slope dH/dx at x = 0.5 to be equal to the slope of the approximation (8.55) at the same x. This gives H(x)»H 0[l + r0(x-i)l.
(8.56)
By itself, (8.56) is not very useful, because (8.55) is nearly as simple and has smaller errors. Its strength lies in the ease of evaluation of certain integrals that involve H(x). For example, an excellent approximation to H(x) can be obtained by writing (8.50) in the form
and substituting (8.56) for H(y) in the integral, which gives
\
(8.57)
This approximation has relative errors smaller than 1% everywhere, which is adequate for virtually all applications. It is interesting to note
8.G. The bidirectional reflectance
213
that if the substitution xln[(l + x)/x]=*2x/(\ + 2x) (Dwight, 1947, eq. 601.4) is made in (8.57), expression (8.55) is obtained. Equation (8.56) also gives useful approximations for the moments of the H functions. Inserting (8.56) into (8.52),
>jTnh[i(ki)\ In particular,
M£)
(8 58a)
'
(8t58b)
8.G.4. Anisotropic scatterers
In the two-stream solution for the diffusive reflectance it was possible to reduce the problem for nonisotropic scatterers to an equivalent problem of isotropic scatterers using the similarity relations. Unfortunately, this is not possible in the solution for the bidirectional reflectance, as emphasized by Sobolev (1975). Although equation (8.37a) can be reduced to an equivalent isotropic form, a nonisotropic scattering function adds an additional term to (8.37b) that cannot be removed except by making the asymmetry factor j8 = 0. Van de Hulst (1974) has shown that similarity relations give excellent results when used in the expression for the bihemispherical or spherical reflectance (see Chapter 10). Unfortunately, although the similarity relations are highly satisfactory for calculating integrated reflectances, they are less so when it is necessary to work with angle-resolved reflectances. The reason for this is that departures from isotropic scattering effectively transfer the scattered radiance from one direction into another. These differences are averaged out in the integrated quantities, but have a much larger effect on the angular distribution of reflectance. Chandrasekhar (1960) has detailed the procedure for finding the exact solution for the bidirectional reflectance of a semiinfinite medium of nonisotropic scatterers and has carried it out for the cases of Rayleigh [p(g) = l + cos 2 g] and first-order Legendre polynomial [p(g) = l +fc1cos g] scattering functions. The solutions are complicated and inconvenient to use. They may be expressed in terms of several functions that satisfy nonlinear integral equations analogous to
214
8. Bidirectional reflectance: semiinfinite medium
(8.50). These functions have been tabulated for certain values of bx and w by Chandrasekhar (1960) and Harris (1957); see also Sobolev (1975), Van de Hulst (1980), and Lenoble (1985). An approximation that is useful if the particle scattering function is not too anisotropic may be obtained by noting that most of the effects of anisotropy are carried by the single-scattering term. As emphasized by Chandrasekhar (1960) and Hansen and Travis (1974), the brighter the surface, the more times the average photon is scattered before emerging from the surface. This tends to average out directional effects in the multiply scattered intensity distribution and causes it to be not too different from the distribution produced by isotropically scattering particles. The dependence of the multiply scattered component of the scattered radiance on pig) is illustrated in Figure 8.9 for the cases where w = l and p(g) = l, 1 + cosg, and f(l + cos 2 g). Although the curve for p(g) = l is too low when pig) = l — cosg, and too high when pig) = 1 + cosg, the shapes of all three curves are very similar. The case when pig) = l ± c o s g is an extreme example, in which the particle scattering functions are highly anisotropic and the large single-scattering albedo exaggerates effects of nonisotropy. When pig) is symmetric, even though it is not isotropic, the multiply scattered Figure 8.9. Multiply scattered component of the radiance scattered into the principal plane from media with w = 1.00 and single-particle scattering functions as shown; e is held constant at 37°, while I varies.
-90
-60
-30
0 i (degrees)
30
60
90
8.G. The bidirectional reflectance
215
component is quite close to that for isotropic scatterers, as illustrated by the curve for Rayleigh scatterers in Figure 8.9. The relative insensitivity of the multiply scattered term to pig) suggests that the solution for isotropic scatterers be used to approximate the multiply scattered contribution to the bidirectional reflectance of a medium of nonisotropic scatterers, while retaining the exact expression for the singly scattered contribution. The exact single-scattering contribution to the radiance at the detector for an arbitrary particle phase function is given by equation (8.35a),
and the contribution of multiple scattering from isotropic scatterers is
Hence, the bidirectional reflectance may be approximated by l],
(8.59)
where H(x) is given by (8.55) or (8.57), depending on the degree of precision required. The adequacy of this approximation clearly depends on the singlescattering albedo and the degree of nonisotropy of the scatterers. The approximation would be poor for a medium consisting of large, weakly absorbing, isolated particles, which have strong diffractive forward scattering. However, in planetary regoliths and laboratory powders the diffractive term is absent, some absorption is invariably present, and the irregular shapes and presence of internal scatterers cause the particle phase functions to be fairly isotropic. For these materials this approximation should be reasonably accurate. When the medium consists of large, well-separated particles, as in a cloud, the diffraction term in p(g) cannot be ignored, and (8.59) will be seriously in error. In this case, Joseph, Wiscombe, and Weinman (1976) suggest treating diffraction as a delta function in the radiative-transfer equation. Lumme and Bowell (1981) have also suggested using the isotropic form for the multiple-scattering term, except that they replace w by w*, where w* is given by the similarity relations (8.30). However, as Sobolev (1975) has emphasized, the similarity relations are valid only for hemispherically averaged fluxes. Figure 8.9 demonstrates that the
216
& Bidirectional reflectance: semiinfinite medium
similarity relations do not necessarily give superior results in this case. When w = 1, w* = w independently of /3; hence, the contribution of the multiple scattering is the same whether w or w* is used. 8.H. The opposition effect 8.H.I. Introduction Some of the changes that occur in the scattering of light from a medium of large particles as their separations change from far apart to close together were discussed in Chapter 7. Here we consider another phenomenon in the same category: the opposition effect, sometimes also called the heiligenschein, hot spot, and bright shadow. The opposition effect is a sharp surge in brightness around zero phase angle, with a typical half-width of 5° to 10°. Its name derives from the fact that the phase angle is zero for solar-system objects at astronomical opposition when the sun, the object, and the earth are aligned. The effect is illustrated in Figures 8.10 and 8.11. The opposition effect is a nearly ubiquitous property of particulate media, including vegetation (Woessner and Hapke, 1987), laboratory powders (Hapke and Van Horn, 1963; Oetking, 1966; Egan and Figure 8.10. Relative brightnesses of an area on the lunar surface [crater Copernicus, from data of Van Diggelen (1965)] and pulverized basalt, illustrating the opposition effect. From Hapke (1968); reprinted with permission of Pergamon Press, Ltd. 1.3 Moon
1.2-
1-5 urn powder
1,1
1.0
0.9
0.8 -10
-5
0
5
Phase angle (degrees)
10
8.H. The opposition effect
217
Hilgeman, 1976; Montgomery and Kohl, 1980), and regoliths of the moon (Gehrels, Coffeen, and Owings, 1964; Whitaker, 1969; Wildey, 1978), Mars (Thorpe, 1978), asteroids (Gehrels et al., 1964; Bowell and Lumme, 1979), and satellites of the outer planets (Brown and Cruikshank, 1983). It is readily observed from an airplane as a bright halo around the shadow of the plane when the shadow falls on vegetation or soil; the halo disappears on pavement. The opposition effect was first discovered by Seeliger (1887,1895) in the light scattered by Saturn's rings, and he also gave the correct explanation of the phenomenon. In a medium in which the particles are large compared with the wavelength, particles near the surface cast shadows on the deeper grains. These shadows are visible at large phase angles, but close to zero phase they are hidden by the objects Figure 8.11. Photograph taken by one of the Apollo 11 astronauts of his own shadow on the lunar surface next to the flagpole. The bright glow around the shadow of the head is the opposition effect. Courtesy of NASA.
218
8. Bidirectional reflectance: semiinfinite medium
that cast them. Thus, the effect may be thought of as being caused by shadow-hiding. Another way of understanding the phenomenon is to think of the interstices between the particles that make up a powder as resembling tunnels through which light penetrates. At large phase angles the interiors of these tunnels visible to the detector are in shadow, the light being blocked by the particles that make up the walls of the tunnels. However, when the phase angle is small, the sides and bottoms of the tunnels are illuminated, resulting in enhanced brightness. The opposition effect is particularly pronounced in fine powders with a mean grain size less than about 20 /im (Hapke and Van Horn, 1963). For particles this fine, the intermolecular adhesive forces that act between the contacting surfaces of two adjacent grains are comparable to the gravitational forces attracting them to the earth. Consequently, only one point of contact is necessary to support a small particle, instead of the minimum of three for the weight of a large particle. Because of this, the microstructures formed by fine cohesive powders can be very porous and intricate, consisting of lacy towers and bridges that Hapke and Van Horn dubbed "fairy castle structures." Because the moon has a strong opposition effect, this showed, even before the Surveyor and Apollo missions, that the lunar regolith is fine-grained and porous. The equation of radiative transfer is inherently incapable of accounting for the opposition effect. However, it has been treated theoretically by several writers in addition to Seeliger, including Bobrov (1962), Hapke (1963, 1986), Irvine (1966), and Lumme and Bowell (1981). Some of the models require a computer to obtain a numerical answer; others give approximate analytic equations that are valid only when the porosity is large. In this section the treatment of Hapke (1986) will be followed, because this gives a convenient, approximate analytic expression whose parameters can be given a straightforward interpretation in terms of physical properties of the medium. The result will be seen to be a modification to the singly scattered, Lommel-Seeliger part of the bidirectional reflectance. 8.H.2. Derivation of the shadow-hiding opposition effect
We begin by considering light that has been scattered only once. In Section 8.G the following expression for the singly scattered radiance
8.H. The opposition effect
219
was derived [equation (8.34)]: lDs(^^g)^J^-f^^(r)p(T,g)ti(r,fi0)tXr^)dr,
(8.60)
where ',(r,Mo)-«~ T / M o (8.61a) is the probability that the incident irradiance / will penetrate to a level r = f™E{z')dz' in the medium, E(z) is the extinction coefficient, E(z) = NE(z)(aQE), NE is the effective particle density, and (aQE) is the average particle extinction cross section. It will be assumed that (aQE), w, and p(g) are independent of T. Similarly, in Section 8.G it was assumed that t§(r,n)-e-*"> (8.61b) is the probability that the light scattered in the direction of the detector by a particle at this level will escape the medium. In this case, (8.60) can be integrated to give the generalized Lommel-Seeliger law,
Note that at zero phase, IDs( e, t ,0) = J-^p(0).
(8.62)
Expressions (8.61) assume that the incident probability tL and exit probability 11 are independent of each other. However, if the phase angle is small, the probabilities are not independent, and (8.61b) overestimates te. To understand this, suppose a ray of light from the source is scattered through phase angle g at some point P located at altitude z. As discussed in Chapter 7, *. is the probability that no particle has its center in a cylinder of radius (aE) whose axis is the ray connecting the source and P, where (aE) =y(crQE)/Tr
(8.63)
is the extinction radius, that is, the radius of an equivalent spherical particle whose cross-sectional area is (crQE). Note that r is the total average number of particles in a cylinder of radius (aE) lying above altitude z. If te were independent of t., then, similarly, te would be the probability that no particle has its center in the cylinder with radius (aE) coaxial with the ray connecting P with the detector. However, portions of the two probability cylinders overlap, as is illustrated
220
8. Bidirectional reflectance: semiinfinite medium
schematically in Figure 8.12, and the probability of extinction by any particles that are in the common volume has been counted twice. Let 2^ be the common volume. In Figure 8.12 the area APBC is the cross section of 2^ in the scattering plane containing P. Correcting for this effect, the product of the incident and escape probabilities can be written t. tt = exp[ - (r/)Lt0 + r/fi - rc)],
(8.64)
rc = / ne(z)d^,
(8.65)
where
and
IDs(e,e,0) = - J^PiO^fj-^dr
= J^p(0),
(8.66)
which is exactly twice (8.62). In effect, the incident ray has preselected a preferential escape path for rays leaving the medium at small phase angles. That is, if a ray is able to penetrate to a given level in the Figure 8.12. Cross section through the scattering plane containing point P. The polygon APBC is the cross section in the scattering plane of the common volume ^ .
8.H. The opposition effect
221
medium and illuminate a point there without being extinguished, then rays scattered from that same point back toward the source will be able to escape without being blocked by any particle. It is often erroneously stated that the opposition effect requires opaque particles. However, the blocking is by extinction, not absorption, and occurs whether the particles are transparent or opaque. It is important to note the clear difference between the scattering properties of a continuous medium and a particulate medium. In a continuous medium, (aE) = 0, so that exponential attenuation occurs along the entire incident and exit paths of the rays, and an opposition effect cannot arise. The common volume 3^ is the intersection of two circular cylinders, of which APBC is the cross section in the scattering plane containing P, and its calculation is mathematically cumbersome (e.g., Irvine, 1966; Goguen, 1981). An approximate value for 2^ may be derived as follows. The area APBC consists of two right triangles with sides (aE) and (aE)cot(g/2), common hypotenuse £ = PC = (aE)csc(g/2), and area (aE)2cot(g/2). Let zx be the projection of £ onto the vertical axis. Then z1 is given by the following system of simultaneous equations:
^ -<—>«« f)"+<—>*. (9 + 9o)2 = £2(tan2 i + tan2 t - 2tan I tan t cos ty)cos g = cos I cos e + sin I sin e cos if/, where ^ is the distance from C to the intersection of the exit ray with the horizontal plane containing C, <^0 is the corresponding distance
222
8. Bidirectional reflectance: semiinfinite medium
for the incident ray, and \fi is the azimuth angle between the projections of the incident and exit rays on the horizontal plane. For small phase angles, c o t ( g / 2 ) » 1, so that
where the positive sign is to be used for ^ and the negative sign for ^ 0 if e> i, and oppositely if I > e. When g is small, each term on the right-hand side of the last two equations is large, but both q, and ^0 are small. Hence, the difference between <j, and ^0 will be small also, and to a sufficient approximation, I f - %\ s zi //* + *i //*o -2
(8.67)
where
I)
or
<M> = i « -
(8.68)
is the average secant of the incident and exit ray paths. Expression (8.67) for zx is a good approximation when g is small. It is poor when g is large and the incident and exit rays are on opposite sides of the normal. However, when g is large, the opposition effect makes only a small contribution to the brightness. Hence, using (8.67) will cause a negligible error in the total radiance for any value of g. At any altitude z' between z and z + zx a horizontal cut through 2^ consists of the common area between two overlapping ellipses. The thickness of this area in the direction perpendicular to the scattering plane is much less sensitive to z' than is the width in the parallel direction. Hence, only a small error will result if 2^ is approximated by a volume of constant thickness u, whose cross section in any plane parallel to the scattering plane is a triangle. This approximation essentially involves replacing the ellipses by overlapping rectangles. We require the area of the triangles to have the same area {aE)2cot(g/2) and projected altitude zx as APBC, and its base is required to lie in the horizontal plane containing P. Then (8.69)
8.H. The opposition effect
223
The portion of 2^ lying above any plane at z' between z and z + zx is Differentiating, dT = (d^ca /dz') dz' = - 2 ^ [ l - ( z ' - z)/zx\
dz'/z,
Thus, the value of rc in (8.64) is
(8.70)
The thickness ^ of the approximated common volume is determined by requiring that TC = T/JU, at g = 0, when (fi) =n = n0 and Zj -*<»; or
{)^
^\y{)
jfNEV){
dz'.
Thus, » =
(8.71)
so that
'-—-^5-jT + * 1 At(»)<«ro J ,>(i - ^ ^ J * -
(8J2)
Integrating (8.72) by parts gives TC = - T / ( / X ) + T 7 ( M ) 5
(8.73)
where
f/Z
+Z
V(z')rfz'.
(8.74)
z
Inserting this result into (8.64) gives (8-75)
224
& Bidirectional reflectance: semiinfinite medium
Thus, the singly scattered radiance can be written ^
.
(8.76)
Equations (8.74) and (8.76) may be readily evaluated for a stepfunction particle density distribution. Let 0 for z > 0, NE = constant for z < 0. Then forz>0,
(0 T'
2
= I T /2T1
\T — TX/2
for - zx < z < 0,
for z
where T^NE{aQE)zly (8.77) and zx is defined by (8.67). Equation (8.76) may be readily evaluated to give
and orf(x) is the error function of argument x. Equation (8.79) is plotted versus y as the solid line in Figure 8.13. It is sharply peaked at y = 0, or g = 0, and decreases rapidly to zero as y and g increase. Although equation (8.79) is analytic, it is inconvenient because most personal computers do not have a subroutine for the error function. To better than 3%, equation (8.79) may be approximated by the function fi(g) = (l + y ) - 1 = ( l + ^ t a n f )
\
(8.81)
where h = \NE{*QE){aE) = - \N{aQE){aE)Hl~^
= ±E{aE), (8.82)
8.H. The opposition effect
225
where > is the filling factor. Expression (8.81) is shown as the dashed line in Figure 8.13. It is much more convenient than (8.79), and the difference between the two curves probably cannot be distinguished in practice.
8.H.3. The angular width of the opposition effect
The angular half-width of the opposition-effect peak is Ag = 2h = NE{aQE){aE) = E{aE). For the step-function particle density distribution this is equivalent to the ratio of the mean particle extinction radius (aE) to the mean extinction length 1/E. In earlier work (Hapke, 1986), equations (8.74) and (8.76) were also evaluated for a density distribution that was exponential with altitude and one that had a hyperbolic-tangent altitude distribution. In these cases it was found that the behavior of B(g) was very similar to that predicted by (8.81), except that the angular-width parameter h refers to the density NE{z) at the altitude where the slant-path optical depth T(Z )/(/*,> =* 1. Equation (8.82) shows that h increases as <j> increases and becomes infinite as <j> -> 1. This simply states that a surface without voids does not have an opposition effect. As
0.8
0.6 B(y)
0.4
0.2
"
226
& Bidirectional reflectance: semiinfinite medium
the peak narrows. Eventually the width of the peak becomes smaller than the angular width of the source or detector as seen from the surface and is then unobservable. The existence of the opposition effect depends on the presence of sharp shadows, and the derivation implicitly assumes that the shadows are infinitely long cylinders. However, when the source of illumination has a finite angular width, the penumbras of the shadows are cones, and a well-defined peak will occur only if one particle has a large probability of being in the penumbra of another particle. An equant, roughly spherical particle of radius a has a cone-shaped penumbra of volume 2TTCI3/30S, where 0S is the angular width of the source. The fraction of the volume occupied by the penumbras is 2Nira3/3ds =
(8.83)
At 1 astronomical unit (AU) from the sun, ds ~ 0.01 and
(8.84)
For a close-packed powder with
H'-/> ln(o,/o,)'
(8.85) l
'
8.H. The opposition effect 227 where al and as are the largest and smallest particle sizes, respectively, in the distribution. Figure 8.14 illustrates the behavior of h when the size distribution is a power law and a / / « 5 = 1,000. If a a i/ s~ l>000 and
4/3^3
A/3/\n( ai /a s ) l/\/2 Note: al and as are the radii of the largest and smallest particles in the distribution, respectively; a is the average particle radius. Figure 8.14. Opposition-effect angular-width parameter for a power-law particle size distribution of the form n(a) a a~v versus v. The ratio of the largest to smallest particles is 1,000.
228
8. Bidirectional reflectance: semiinfinite medium 8.HA. The amplitude of the opposition effect
8.H.4.a. The opposition effect in multiply scattered light
It is easy to show that the shadow-hiding effect is negligible for the multiply scattered component. The enhanced brightness is caused by the overlap of the incident and emergent probability cylinders associated with a ray scattered from a single point within the medium. If the ray is scattered between two or more particles before emerging, the probability that the incident and emergent cylinders will overlap is very small. Esposito (1979) carried out a numerical calculation of the opposition effect including doubly scattered light and found a negligible addition to the reflectance. 8.H.4.b. Effects of finite particle size
Theoretically, the opposition effect should increase the singly scattered component of the radiance by a factor of exactly 2 at zero phase, that is, 5(0) = 1. The derivation of the opposition effect assumed that the ray was scattered by a point; that is, the finite size of the particle was neglected. However, suppose a ray is scattered only once by a particle in the medium, but the point at which it leaves the particle is different from the point at which it enters. Then the amount of overlap between the probability cylinders is less than that calculated in equations (8.78) and (8.81), and the amplitude of the opposition effect is decreased. If the ray is scattered by specular reflection directly from the particle surface, or after penetrating only a short distance into the particle and being scattered back out by subsurface scatterers, then the overlap is large, and the amplitude of the opposition effect will be close to its theoretical value of 1. However, the refracted rays and the internally reflected, backscattered rays may enter and leave the particle at points that are separated by up to nearly the entire particle diameter, and for this light the opposition effect is negligible. In order to take account of these effects, B(g) in (8.81) may be multiplied by an empirical factor Bo. For a true, shadow-hiding opposition effect, 0 < BQ < 1, and Bo is the ratio of the light scattered from near the illuminated surface of the particle to the total amount of light scattered at zero phase: B0 = S(0)/wp(Q), (8.86) where 5(0) is the light scattered at or close to the part of the surface
8.H. The opposition effect
229
of the particle facing the source, and wp(0) is the total amount of light scattered by the particle at zero phase. A lower limit to 5(0) occurs when the particle is opaque. Then 5(0) is the specular component of the particle scattering function,
the Fresnel reflection coefficient at normal incidence. However, subsurface scattering (Chapter 6) will increase 5(0) above this value for most particles except true opaques. 8.H.4.C. Coherent backscatter
Because particles with size parameter X < 1 do not have well-defined shadows, a shadow-hiding opposition effect will not be observed on a medium made up of such particles. However, when X ~ 1 an entirely different phenomenon can cause a surge in brightness at small phase angles. The phenomenon is known as coherent backscatter or weak photon localization. At optical wavelengths it has been investigated both experimentally and theoretically by several researchers, including Kuga and Ishimaru (1984) and Wolf and Maret (1985), and has been discussed in detail by MacKintosh and John (1988) and Hapke (1990). It was independently suggested as a cause of the opposition effect by Shkuratov (1988) and Muinonen (1990). The term weak photon localization comes from the fact that the phenomenon occurs in the transition region between physical optics, where the wave nature of electromagnetic radiation must be taken into account, and geometric optics, where the light may be considered as localization and the phase paths may be described by rays. Briefly, two partial waves associated with the same incident wave front may travel the same multiply scattered path between particles of the medium, but in opposite directions. If the points at which the two partial waves emerge from the medium are not separated by a distance large compared with the wavelength, then upon emerging from the medium the waves will combine coherently. The waves will be in phase at exactly zero phase angle. Because the scattered power is determined by squaring the total amplitude, at zero phase the power of the two combined portions of the waves is four times that of a partial wave, instead of twice. The angular width of the coherent-
230
8. Bidirectional reflectance: semiinfinite medium
backscatter peak is approximately Ag-A/27r^,
(8.88)
where J? is the photon mean free path in the medium. The coherent-backscatter opposition effect can be distinguished observationally from the shadow-hiding opposition effect in two ways: (1) The width of the coherent-backscatter peak increases linearly with A, whereas that of the shadow-hiding peak is independent of A. (2) Coherent backscatter depends on multiply scattered light; therefore, the amplitude of the peak relative to the continuum increases as the reflectance increases. Conversely, shadow hiding is important only for singly scattered light; therefore, the height of the peak relative to the continuum decreases as the reflectance increases. Typical coherent-backscatter angular widths observed in the laboratory are of the order of 0.5°, whereas the widths of the opposition effects observed on inner bodies of the solar system are an order of magnitude larger. Hence, the latter probably are due mainly to shadow hiding, although coherent backscatter may make an important contribution. However, much narrower surges have been measured on icy satellites in the outer solar system (Brown and Cruikshank, 1983; Domingue et al., 1991), and these may be due mainly to coherentbackscatter. The phenomenon also appears to be important in the scattering of radar waves from icy satellites (Hapke, 1990; Hapke and Blewett, 1991). 8.H.4.d. Other contributions to an opposition effect
There are several other phenomena that may also cause a surge in brightness at small phase angles. These include the following: (1) An intrinsic single-particle opposition effect: The particles making up the medium may consist of complex porous agglomerates of smaller particles that are still large compared with the wavelength. This will cause an additional shadow-hiding opposition effect that will be intrinsic to the particle phase function p(g). (2) Glory: If the particles are spherical they will have a glory. (3) Internal reflections: Transparent particles with plane surfaces that meet at angles close to right angles, such as minerals with a cubic structure, may cause a brightness surge by a corner-reflector effect, in which rays are triply totally internally reflected back toward the source (Trowbridge, 1978; Muinonen et al., 1989).
231
8.H. The opposition effect Figure 8.15a. Bidirectional reflectance in the principal plane of a medium of scatterers with e = 60°, w = 0.99, p(g) = 1 + 0.5 cos g, Bo = 1.00, and h = 0.06. The solid line is calculated from (8.89). The dashed line is Lambert's law. 0.3
0.25 0.2
0.15 0.1
0.05
-90
-60
-30
0 i (degrees)
30
60
90
Figure 8.15b. Bidirectional reflectance in the principal plane of a medium of scatterers with e = 60°, w = 0.19, p(g) = 1 + 0.5cos g, Bo = 1.00, and h = 0.06. The solid line is calculated from (8.89). The dashed line is Lambert's law. 0.025 W = 0.19 0.02
0.015
0.01
0.005
-90
-60
-30
0 i (degrees)
30
60
90
232
8. Bidirectional reflectance: semiinfinite medium However, the lack of optical perfection of natural particles probably makes this effect of minor importance for most media (Hapke and Williams, 1989). These various phenomena may be large enough to increase the opposition surge by more than a factor of 2. This possibility may be taken into account by allowing BQ to be greater than 1. Figure 8.16. Bidirectional reflectance versus I for a powder of size < 37 ium made from the cobalt glass whose absorption spectrum is shown in Figure 6.16. The detector views the powder at e = 30°, while I is varied in the principal plane. The dots show the reflectance measured at five wavelengths in the ultraviolet (U), blue (B), green (G), red (R), and infrared (I). The lines show equation (8.89) fitted to the data. From Hapke and Wells (1981); copyright by the American Geophysical Union.
90
30
0
30
60
90
8.H. The opposition effect
233
8.H.5. The bidirectional reflectance including the opposition effect
Combining the results of Sections 8.G and 8.H, we may write, for the bidirectional reflectance,
(8.89) where B(g)~
(l//*)tan(g/2)' h is given by (8.82), and H{x) by (8.55) or (8.57). Equation (8.89) is the basis for the analytic approximations to a number of quantities of interest in reflectance spectroscopy that will be derived in the remainder of this book. The ability of equation (8.89) to describe the bidirectional reflectances of a wide variety of particulate surfaces has been tested in several papers (Hapke and Wells, 1981; Pinty, Verstraete, and Dickinson, 1989; Clark, Kierein, and Swayze, 1993). The general character of this bidirectional-reflectance function is illustrated in Figure 8.15, which plots r( i, e, g) for surfaces of high and low albedo. Figure 8.17. Relative brightness profile along the equator of Venus as a function of longitude. The crosses show data measured by the Manner 10 spacecraft. The line shows equation (8.89). From Hapke and Wells (1981); copyright by the American Geophysical Union.
Position on Disk
234
8. Bidirectional reflectance: semiinflnite medium Figure 8.18. Comparison (crosses) between the reflectances in visible light of the bare soil in a plowed field predicted by the theory and measured in various geometries. If the predicted and measured values agreed exactly, the crosses would fall on the straight line. Reprinted by permission of the publisher from Pinty et al. (1989). Copyright 1989 by Elsevier Science Publishing Co., Inc.
0.1 0.2 0.3 Measured Bidirectional Reflectances
0.4
Figure 8.19. Theoretical bidirectional-reflectance function fitted to measurements on various solar-system bodies near opposition. From Hapke (1986), courtesy of Academic Press, Inc.
Am
8.H. The opposition effect
235
Also shown for comparison is the Lambert reflectance function. Comparisons with experimental data are shown in Figures 8.16, 8.17, and 8.18. The ability of (8.90) to describe the opposition effect is shown in Figure 8.19. For example, the value of h for the moon is 0.05. The size distribution of lunar regolith approximates a power law with exponent — 4, From this it may be calculated that the filling factor is 0.41, which agrees with the measured value of 0.27-0.65. More examples are given in Bowell et al. (1989). However, it must be emphasized that the coherent-backscatter opposition effect has only recently been recognized, and its importance in planetary remote sensing has not been adequately assessed, particularly in high-albedo objects.
9 The bidirectional reflectance in other geometries
9.A. Introduction
Chapter 8 treated the bidirectional reflectance of an optically thick, plane-parallel participate medium in which the particles were randomly oriented and could be regarded as embedded in a vacuum. In this chapter we will discuss the effects on the reflectance when each of these restrictions is removed. 9.B. Diffuse reflectance from a medium with a specularly reflecting surface
The upper surfaces of many particulate materials may be sufficiently smooth on a scale comparable to the wavelength that light is scattered both quasi-specularly from the surface and diffusely from below the surface. The specular component is known as regular reflection. Because a surface effectively becomes more optically smooth at large angles of incidence (Chapter 6), the regular component may become especially important at large phase angles. The most familiar example of the combination of diffuse reflection and regular reflection is water containing suspended solids, as in rivers, lakes, and oceans. If a body of water is examined in the geometry for specular reflection from the surface, a bright glare is seen, which is the reflected image of the sun. However, if the same body is examined in an off-specular configuration, it looks dark and may be colored blue, brown, or green, depending on the nature of the 236
9.B. Diffuse reflectance with a specularly reflecting surface
237
suspended solids. Combinations of specular and diffuse reflectances are encountered in many other materials, including glossy paints, glazed ceramic tiles, glazed paper, polished wood, and leaves with waxy surfaces. In some commercial reflectance spectrometers the sample is held vertically or upside down, necessitating the use of a specularly reflecting cover glass if the material is a powder. A first-order expression for the specular-diffuse, bidirectional reflectance rsd(i,e,il/) of such types of surfaces may be obtained by simply combining the specular- and bidirectional-reflectance laws, after correcting for transmission through the specular surface:
R(6)]r(i'9e'9t).
(9.1)
Here R(x) is the Fresnel reflection coefficient averaged over the two directions of polarization; 8(x) is the delta function, r(i\ e,il/) is the bidirectional reflectance of the scattering medium under the smooth surface, and I1 and e! are given by the Snell relations, sin I = n sin i' and sin e = n sin e', where n is the index of refraction. Equation (9.1) is an oversimplification in several respects. First, it assumes that the regular component remains collimated after reflection, whereas most natural surfaces have some large-scale roughness structure, in which to first order the surface may be treated as made of up smooth facets tilted in a variety of directions. The roughness spreads the specularly reflected light out into an elliptical cone whose angular half-width in the principal plane is approximately equal to twice the mean surface tilt. Second, at small phase angles the radiance reflected from a rough surface is maximum at the specular angle, but at large angles of incidence and reflection the maximum intensity is skewed slightly toward phase angles larger than specular. The reason for this is that the Fresnel reflection coefficients are rising rapidly and nonlinearly at large angles, so that light reflected from surface facets tilted slightly away from the source makes a much larger contribution to the radiance than that from the facets tilted toward the source. Third, equation (9.1) ignores the diffuse light that is multiply internally scattered between the surface and the scatterers. However, the lowest-order term of this component is of order [1 - R]2Rr2 and, because R is usually small, can be ignored in most applications.
238
9. The bidirectional reflectance in other geometries
A particularly interesting application of combined specular and diffuse reflectances is the scattering of radar waves from planetary surfaces. If an inner planet, such as the moon, is illuminated by radar, most of the power is scattered back to the antenna by specular reflection from the surface, which is relatively smooth on the scale of radar wavelengths (Evans, 1962; Evans and Hagfors, 1971). However, the return also displays a small diffuse background (Figure 9.1) whose size, relative to the specular component, depends on the planet and the wavelength. The nature of this diffuse component is uncertain. Most workers in the field of radar astronomy attribute it to specular reflection from surface facets tilted toward the antenna on exceptionally rough portions of the planet. However, if the interior of the regolith is sufficiently inhomogeneous, then some or all of the diffuse component must come from radiation that has been transmitted through the surface and diffusely scattered by the inhomogeneities (Thompson et al., 1970; Pollack and Whitehill, 1972). Certain satellites of the outer planets have strong diffuse components (Ostro, 1982), and it has been suggested that subsurface scattering is the dominant process in the reflectance of radar waves from these bodies (Ostro and Shoemaker, 1990; Hapke, 1990). If the diffuse radar component is due to subsurface scattering, as is very likely the case, then equation (9.1) should apply. Monostatic Figure 9.1. Typical radar signals received from solar-system objects in circularly polarized light. The inner bodies, such as the moon, have a strong specular component and weak diffuse component, whereas the situation is reversed for outer satellites, such as Europa. The terms "expected" and "unexpected" refer to the sense of circular polarization that would be specularly reflected from a plane surface with circularly polarized radiation incident. From Ostro (1982); copyright 1982; University of Arizona Press. 0.5
- Expected polarization Unexpected polarization Europa
0.25
Inner Planets
0
-
1
+1
Relative Doppler Frequency Shift
0
-1
239
9.C. Oriented scatterers: applications to vegetation
radar observes an area on a planet at g = 0 and I = e = ft, where ft is the angular distance between the area and the sub-radar point subtended at the center of the planet (Figure 9.2). Then the combined specular-diffuse reflectance per unit area is
rsd(ft,ft,0)cosft = R(ft)8(ft) +
[l-R(ft)]2r(ft',ft',0)cosft', (9.2)
where sin ft = n sin ft'. In practice, the regular component is not exactly a 5 function, but has a finite angular width, which can be related to the mean tilt of the surface of the planet (Hagfors, 1964; Evans and Hagfors, 1971; Simpson and Tyler, 1982). If the sizes of the subsurface inhomogeneities are comparable to the wavelength, then coherent backscatter will occur. This phenomenon is important in radar scattering by outer satellites (Hapke, 1990). 9.C. Oriented scatterers: applications to vegetation canopies
Some media consist of asymmetric particles with orientations in a preferred direction. An important example is vegetation, in which the "particles" are leaves, which are flattened structures whose faces are Figure 9.2. Geometry of radar scattering from a planet.
Planet Radar Wave
240
9. The bidirectional reflectance in other geometries
sometimes oriented with respect to the direction of sunlight or prevailing wind. A second example is a sediment in which the deposition process has preferentially oriented the particles. A third example occurs when a powder is subjected to such a high pressure that the grains are reoriented or even deformed so that they tend to lie with flat sides parallel to the surface of the medium. Such a medium will have a strong specular component. Kortum (1969) gives several examples of pressure-induced specular reflection. Blevin and Brown (1967) noted that colloidal suspensions had higher reflectances at lower porosities than at higher. Hapke and Wells (1981) found a similar effect: A powder consisting of fine silicate glass particles was systematically brighter and more forward-scattering when closely packed than when loosely sifted. These results could be explained by preferential orientation. If the particles are not randomly oriented, all of the scattering parameters defined in Chapter 7 will depend on the mean orientations of the particles with respect to the directions of incidence and emergence. In general, preferentially orienting the particles will completely change the angular pattern of the singly scattered component of the reflectance. Instead of the bidirectional reflectance being described by (8.89), the single-scattering term preserves the angular-scattering properties of the individual scatterers. For example, a semiinfinite medium consisting of parallel mirrors would have a specular, rather than diffuse, single-scattering law. It is only the multiply scattered component that tends to diffuse and randomize the light scattered by the individual particles. Scattering by vegetation canopies is a major topic in terrestrial remote sensing. It has been discussed by many authors using a variety of approaches, including Suits (1972), Kimes and Kerchner (1982), Gerstl and Zardecki (1985b), Otterman (1983), Camillo (1987), Ross and Marshak (1984), Verstraete, Pinty, and Dickinson (1990), Pinty et al. (1990), and Dickinson, Pinty, and Verstraete (1990). See also Smith (1983). Woessner and Hapke (1987) found that equation (8.89), which was derived for soils, gave a reasonably good fit to observational data on clover (Figure 9.3). Verstraete et al. (1990) and Pinty et al. (1990) have extended the model on which (8.89) is based to the case of preferentially oriented scatterers. Although their model was derived specifi-
9.C. Oriented scatterers: applications to vegetation
241
cally for applications to vegetation, a similar analysis applies to any medium in which the scatterers are not randomly oriented. Suppose the particles are not equant, but are relatively flattened disks with mean cross-sectional area a and with a preferred orientation. For instance, the disk could represent leaves or portions of leaves. Then, in the terms involving the incident light in the radiativetransfer equation, a must be replaced by at, where aL is the average of the projected cross-sectional areas of the particles onto a plane perpendicular to the direction of incidence. Let Mi-oi/cr. (93) Similarly, in the terms involving the emerging rays, a must be replaced by er6, the mean projected area of the scatterers onto a plane perpendicular to the direction to the detector. Let A*. = * ; / * •
(9.4)
In vegetation, fii and \xt involve the leaf droop, which is a measure of the tilt of the average leaf surface from horizontal. The multiple-scattering contribution to the scattered radiance inside the medium is much less sensitive to both the particle angularFigure 9.3. Scattering by clover. The crosses are values measured in the principal plane as a function of e with I — 0; the line is the theoretical bidirectional-reflectance function, equation (8.89). Reprinted by permission of the publisher from Woessner and Hapke (1987). Copyright 1987 by Elsevier Science Publishing Co., Inc. i
0.1
.08
1 cc
i
I
i
i
1
i
i
|
i
i
|
'
•
I
-
'
.06
1
CD
.04
* ^
.02
0
i
90
i
1
60
<
i
1
i
1
,
30
I
1
30 e (degrees)
1
1
1
60
.
1
90
242
9. The bidirectional reflectance in other geometries
scattering function and orientation. Hence, the same simplifying approximation may be made as in the derivation of (8.89), namely, that the radiance multiply scattered within the medium is similar to that which would occur if the particles were randomly oriented, isotropic scatterers. Then it is straightforward to show that these modifications result in an equation almost identical with (8.89), except that /JL0 and /x, are replaced by JLL 0 /^. and //,///, e, respectively. With these modifications, the generalization of the bidirectionalreflectance model to the case of oriented scatterers, including vegetation canopies, is
}
(9.5) As with other particulate media, a vegetation canopy exhibits a strong opposition effect, which is often called the hot spot The hot spot may be described by an equation of the form of (8.69), except that the interpretation of some of the parameters is changed. Leaves have complex structures that act as internal scatterers. Hence, much of the light backscattered at small phase angles comes from the parts of the leaf surfaces close to the points at which the rays entered, so that the opposition-effect amplitude parameter Bo should not be very different from 1. In the angular-width parameter h = [equation (8.82)], the leaf filling factor is sufficiently NE(crQE)(aE)/2 low that NE ^ N. Because the leaves are large, QE = 1. The size distribution of the leaves is fairly narrow, so that (crQE) = o-, aE = (O-/TT) 1 / 2 , and (o-QE)(aE) = O - 3 / 2 / T T 1 / 2 ; however, a must be corrected for leaf orientation. Within the opposition peak, /i. = /x,e; hence, to a sufficient approximation, a may be replaced by 07Z, where 2. Thus,
Note that the angular width of the hot spot decreases as the leaves get farther apart and that h also depends on the leaf droop.
9.D. Layered media 243 The leaf-area index LAI(z) at an altitude z is defined as the total fraction of the area above that altitude occupied by leaves:
rz',
(9.7)
where fih is the cosine of the angle between the perpendicular to a and the vertical. Hence, for horizontally oriented leaves the leaf-area index equals the optical depth: L A I ( Z ) = T(Z). If the integrand is approximately constant, LAI(z) = Na/jihz, where z is the depth below the top of the vegetation canopy. Thus, the hot-spot angular-width parameter is proportional to the leaf-area index: /i = LAI(Az), (9.8a) where 1/2-3/2
f ^
<9-8b>
For example, if the leaves all lie horizontally and the canopy is illuminated vertically, h is equal to the leaf-area index of a layer whose thickness Az is approximately one-quarter of the mean leaf diameter. Verstraete et al. (1990) have suggested an alternative expression to describe the hot spot. However, it is not clear that their expression is more accurate than (8.90). Before leaving this topic, it must be emphasized that these equations are only a beginning. The general problem of the scattering of radiation by vegetation is formidable indeed, and it requires treating at least four layers of materials: the atmosphere, the vegetation canopy, the floor litter, and the soil. Additionally, the plants may not be randomly positioned, but aligned in rows. The treatment of these problems is beyond the scope of this book. 9.D. Layered media 9.D.I. Introduction
There are many instances in remote sensing where a surface is overlain by a layer of a different kind of material. Examples are an atmosphere or haze above a planetary regolith, dust or frost on a planetary surface, and a vegetation canopy over soil.
244
9. The bidirectional reflectance in other geometries
To introduce the subject of scattering by layered media, expressions for the diffusive reflectance and transmittance of several different configurations of layers will be derived. The mathematical simplicity of these expressions makes them extremely useful for quantitative, first-order estimates of situations often encountered in planetary and remote-sensing applications. The solutions will be obtained for isotropic scatterers, but can be easily generalized to include nonisotropic scatterers by using the similarity relations, equations (8.30). Finally, the bidirectional reflectance of a two-layer medium will be derived.
9.D.2. The diffusive reflectance and transmittance of layers
9.D.2.a. The radiance in a layer offinitethickness We begin by considering a layer consisting of isotropically scattering particles with effective density NE9 cross-sectional area
TO= J
— 00
E = NEaQE is the extinction coefficient, and z is the altitude. Obviously, in order that r 0 be finite, NE must approach zero above and below certain altitudes that may be described as the top and bottom of the layer, respectively. Let the albedo factor be y = \ / 1 - H > , SO that if the layer were infinitely thick it would have a diffusive reflectance r%, of
In the diffusive two-stream solution to the equation of radiative transfer, the radiance is governed by equations (8.17), which have the
9.D. Layered media
245
general solution of the form given by equations (8.23), y)e2yT],
(9.10a) (9.10b)
where Ix and / 2 are, respectively, the upward- and downward-going radiances in the layer, r is the optical depth, and sf and & are constants to be determined by the boundary conditions. Various geometries may be investigated by imposing appropriate boundary conditions on equations (9.10). The reflectances and transmittances of layered media of nonisotropic scatterers may be found from the solutions for isotropic scatterers by replacing w and r by w* and r*, respectively, using equations (8.30) or (8.32). 9.D.2.b. Two-layer media
Consider the case where the layer lies on top of a second material that is infinitely thick and has diffusive reflectance r^. The boundary conditions are that the radiance must be continuous across the top and bottom surfaces of the upper layer. At the top of this layer, this requires that / 2 ( 0 ) - | [ j / ( l + y) + ^ ( l - y ) ] - / 0 .
(9.11)
The interface between the two layers is at r = r 0 . At this interface the upwelling radiance must be equal to the fraction of the downwelling radiance reflected back up from the lower layer: or •&(l + y)e2yTo] y)e~2yT° + ^ ( 1 - y)e2yT°].
(9.12)
Solving equations (9.11) and (9.12) for $f and & gives 1 + y (1 -
T^7(i-r
r
) + r (r -r )e~A^'
(
}
246
9. The bidirectional reflectance in other geometries
The diffusive reflectance r 0 of the two layers is the fraction of the incident power emerging from the upper layer, r
°-
TT/0
"
sf{\ - y) + &(1 + y) 2/ 0
Inserting the foregoing expressions for j& and & gives
(9.14) Let us explore some of the properties of this solution for r0. The radiance in the upper layer and the degree to which the reflectance of the lower layer influences the total reflectance are governed by the quantity 4yr 0 . As T 0 -»O°, ^0"*%? the diffusive reflectance of the upper layer, but the layer may be considered to be optically thick when yr0 > 1. When r 0 -»0, r 0 -> r#, the diffusive reflectance of the lower layer. The downwelling radiance at the base of the upper layer is Ioto, where t0 is the diffusive transmittance of the upper layer, given by y)e ~2yT° Substituting the expressions for s/ and & gives
(9.15) Values of the transmittance range from tQ = 1 at T 0 = 0 to t0 = 0 as Equation (9.15) is useful for estimating the radiance at the interface between two media, such as at the base of an atmosphere or at the bottom of a snow deposit. An interesting situation occurs when there is negligible absorption in the upper layer. This might approximate the case of clouds or snow covering the surface of a planet. When w — 1, y
T 0 -»OO.
9.D. Layered media
247
(9.14) becomes
To first order in y, r<% ^ 1 - 2 y , and 1 — r£ = 4 y / ( l + y) 2 = 4y, so that (9.16) Similarly, (9.17) When r 0 = 0, r0 = rg, and t0 = 1; when T0 » 1, r 0 ~ 1 and tQ = 0; when r* = 0, r0 = r 0 / ( I + T 0 ) and t0 = 1/(1 + r 0 ). Although the diffusive-reflectance model is not exact, nevertheless it is capable of giving answers that are sufficiently accurate for many applications. This is illustrated by Figure 9.4, which shows the two-layer diffusive-reflectance model, equation (9.14), fitted to a series of measurements by Wells, Veverka, and Thomas (1984). Thin layers of fine Figure 9.4. Reflectances of thin, uniform layers of fine volcanic ash on bright and dark substrates plotted against the thicknesses of the layers. The points are values measured by Wells et al. (1984); the lines are equation (9.14) with the particle sizes that give the best fits. Mauna Kea Volcanic Ash White substrate
Theory (D = 3.3n.m)
0.5
thickness (mg/cm
1.5
248
9. The bidirectional reflectance in other geometries
volcanic dust 1-5 fim in diameter were deposited on top of surfaces covered with white paint and black paint. The bidirectional reflectance K6°,0o ,6°) was measured as a function of dust thickness in units of mass per unit area, J? mg/cm 2 . Now, for a uniform upper layer of thickness S?, r 0 = NEaQEJ?. Assume that NE is not very different from N. Then for equant particles of diameter D » A and density p, Q E = 1, J? = N(TT/6)D3P&, and a=(ir/4)D2. Hence, T 0 ==f(^f/pD). The only adjustable parameter in these fits is the mean dust-particle diameter. The theoretical sizes for best fits were 2.0-3.5 ^m, which are acceptably within the measured size range. The effects of layering on spectra are discussed in Chapter 11. 9.D.2.C. The reflectance and transmittance of an isolated layer
The reflectance and transmittance of a layer suspended in space or lying on a material of very low albedo may be found from (9.14) and (9.15) by setting r# = 0. This gives
'•-''wjX-
(9 18a)
-
The total fraction of incident light scattered by the layer is the sum of the reflectance and transmittance,
9.D.2.d. Reflectance and transmittance of an absorbing slab containing embedded scatterers
In the discussion of the effects of internal scatterers on the efficiencies of a large irregular particle (Chapter 6), the properties of a slab containing embedded scatterers were needed. This problem will now be solved by using the two-stream method to calculate the diffusive reflectance and transmittance of an absorbing slab containing distributed scatterers.
9.D. Layered media
249
The problem to be solved is the following: Diffuse radiance Io is uniformly incident from the hemisphere above a plane slab of material having real refractive index n, absorption coefficient a, and distributed internal scattering coefficient 6. The thickness of the slab is J?. The reflection coefficient of the surface of the slab for diffuse light externally incident is Se, and for diffuse light internally incident, 5,-, where Se and St are functions of m and were defined in Chapter 6. The light is assumed to be scattered isotropically by the distributed internal scatterers and by the surfaces of the slab. Inside the slab, K = a, S = 5, and E = a + 5; hence, w = S/E = 4). From equations (8.23), the 4 / ( a + 6) and y = }/l-w =Ja/(a+ general solution for the radiation field inside the slab is [^(i
y
)
e
hi?) = M^C 1 + y)e'2yT
+&(1 +y ) e ] ,
(9.20a)
+ •#(! - r)e 2yT ] >
(9.20b)
where r = — Ez = — ( a + a)z. Let the upper surface of the slab be at z = 0, and the lower surface at z = — Sf. The optical thickness of the slab is r 0 = (a + 4)-S*, and 2yr0 = 2^/a(a + 4 ) ^ . The boundary conditions are as follows: At the lower surface of the slab, T = TQ, a fraction S, of the downgoing radiance is reflected by the interface back into the upward direction. That is,
or
At the upper surface, a fraction (1 — Se) of the incident radiance / 0 is transmitted through the surface into the slab, and a fraction St of the upgoing radiance is reflected by the surface back into the downward direction inside the slab. Thus, 72(0) = ( l - 5 j / 0 + 5 / / 1 (0), or 2
250
9. The bidirectional reflectance in other geometries
Solving these equations for sf and & gives 1+
y ( 1 - 5,.r,.)2-(5, - r,) 2 exp(-4Va(a + i)&)
°' (9.21a)
s
2
1+
Y ( 1 - S,r,) 2 -(S, - r,.) 2 exp(-4 1 /a(a + (9.21b)
where
is the equivalent diffusive reflectance of a medium with K = a, S = 1, and £ = a + i. The external radiant power incident per unit area on the upper surface of the slab is ir/ 0 . The power per unit area emerging from the upper surface is Tr[SeI0+(l — 5,)7^0)], and that from the lower surface is ir(l — 5 , ) / 2 ( T 0 ) . Hence, the reflectance of the slab is
C (
eK1
" S | ) (l-W-(5 l -r l ) a «p(-4VS(^n
(9.22)
and the transmittance is
(1 - 5,r,.)exp( -2Va(q + i ) ^ ) + ro(5, - r,)exp( (1 - 5 ; r,) 2 - (5, - r,)2exp( -4^a(a +
(9.23)
9.D. Layered media
251
The total fraction of incident light scattered by the slab is
X
r, + e x p ( - : *—
If the slab is taken to be a "particle," then r 0 + 1 0 = Qs, the nondiffractive part of the scattering efficiency. After a little algebra, the last equation may be written in the form >
( 9 - 24a )
r, +exp(-2\/a:(a+ { 7, -(— V V ' , 0 =—
(9.24b)
Qs = Se + (1 - Se) \_SQ where
which was the equation used in Chapter 6. If the slab is optically thick, 2ja(a+ 6)5? » 1, v =S 1 (1 S )(1 S)
*
(9 25)
If we allow r 0 -»oo ? equation (9.25) also represents the diffusive reflectance of an optically thick layer of powder underneath a cover glass. If the internal scatterers are not isotropic, but can be described by the hemispherical asymmetry factor /3, the reflectance and transmittance of the slab can be found from the similarity equations (8.30) by substituting w* = [(1 - j6)/(l - pw)]w and 3?* = (1 - )3)J? for w and J?, respectively.
9.D.3. The bidirectional reflectance of a two-layer medium
Although the diffusive reflectance is convenient for estimating the radiance in a double-layer geometry, a theory of bidirectional reflectance is necessary for more precise calculations. In this section we
252
9. The bidirectional reflectance in other geometries
will derive an approximate expression for the combined bidirectional reflectance of an optically thin layer overlying a thick medium. Consider an optically thin layer that contains particles of singlescattering albedo w% and angular phase function p^(g), and with extinction coefficient E#, lying on top of a second material of very large thickness, and containing particles of single-scattering albedo w# and phase function p&ig), and with extinction coefficient E#. See Figure 9.5. The optical thickness of the top layer is T 0 , and z0 is the altitude of the interface. The medium is illuminated from above by a distant collimated source of irradiance / incident from a direction making an angle I with the vertical, and viewed by a distant detector from a direction making an angle e with the vertical. Using the same procedure as in Chapter 8, an exact expression will be found for the singly scattered radiance for particles of arbitrary phase function. An approximate expression for the multiply scattered radiance will be found using the two-stream approximation with collimated source, and assuming that the particles scatter isotropically. From equation (8.8), the radiance at the detector can be written Figure 9.5. Schematic diagram of the two-layer bidirectional-reflectance problem.
p(g)
= p,
9.D. Layered media
253
in the form
xf p(T,a',£l)I(T,nf)dn']e-T/lxdT. J
4ir
(9.26)
J
For this problem, , O) =
/^P(T,
*)«-'"*,
(9.27a)
where \
(9.27b)
(T>T0).
Using the same assumptions and procedures as in the derivation of the bidirectional reflectance of a single, optically thick layer, the integral inside the brackets is approximated by ),
(9.28)
where
I
(T
w<%e
T
^0<%e
r Br^e
^T
\
<^ r0)
J**e-2y*T + a:re2y*T + 9'jre-T/''° ( T > T 0 ) . ' In addition, the solution requires the quantity 1 The constants ^ and ^ can be found by inserting (9.29) into the differential equation for cp{r) and equating coefficients of e~T//x°. The remaining four constants are found from the boundary conditions as follows: (1) cp{r) must remain finite as T-><X>. This requires that && = 0. (2) The upward- and downward-going radiances must be continuous across the interface at T 0 . This is equivalent to requiring that cp and A
254
9. The bidirectional reflectance in other geometries
where rs and rm are, respectively, the singly and multiply scattered components of the reflectance and are given by the following expressions:
(9.31a)
r (i e e) = — I
—
fl-
(1/ll+2y )To
e-
* ]
(9.31b) (9.32a) (9.32b)
(9.32c) - - [(1 - y * ) / ( i + y * ) ] ^ - [(i + i / 2 / t o ) / ( i +
(9.32d) £ ^ 13"' 0 ^
(9.32e) The opposition effect is still given by the usual expression Bn
(9.33)
9.E. Spherical geometry
255
provided that the parameters Bo and h refer to their values at the slant-path r = 1 level, which may be in either the upper or lower layer, depending on the values of T 0 and I. Note that equations (9.31)-(9.33) for r(i, e,g) include the specific bidirectional scattering properties of both the lower and upper layers, all orders of multiple scattering within both the lower and upper layers, and all orders of multiple scattering between the two layers. These expressions should be useful for calculating the combined bidirectional reflectance of such systems as a hazy atmosphere over a planetary regolith and the reflectance of a vegetation canopy over soil. For the latter application, the fact that the scatterers may not be randomly oriented may have to be taken into account, as discussed in Section 9.C. 9.E. Spherical geometry 9.E.I. Single-scattering theory In a soil, the thickness of the layer from which the light is scattered is so small relative to the radius of the body that the surface may be treated as plane-parallel right up to the limb or terminator. However, when the light is scattered from a cloud or haze layer it is often necessary to take the curvature of the surface into account. Light scattering in a spherical atmosphere was first treated by Chapman (1931), and by many others since then (e.g., Rozenberg, 1966). The spherical geometry makes the radiative-transfer problem even more intractable than in the case of plane geometry, but numerical solutions have been given by several authors, including Adams and Kattawar (1978) and Kattawar (1979). However, it is possible to obtain approximate, analytic solutions in the interesting case in which the scale height of the scatterers in the atmosphere is small compared with the radius of the body (Wallach and Hapke, 1985). Assume that the scattering layer is optically thick and that the density of scatterers N(&) at a radial distance & from the center of the planet is given by
(
go
- - * -
c& \
1
) ,
(9-34)
256
9. The bidirectional reflectance in other geometries
where <%* is the scale height of the aerosols, and 31x is the radius at which the optical depth (looking radially) is 1. Assume that w and pig) of the scatterers are independent of altitude. Let c^o be the radius of some arbitrary reference surface from which the light is interpreted as being scattered. This surface may, for example, be that of a sphere that is tangent to the line of sight of the detector, or the radius at which the optical depth is 1. At the point where the line of sight of the detector intersects the sphere ^ 0 , the local angle of emergence between the radius vector and the direction to the detector is e0, and the angle between the radius and the incident ray is i0. The geometry is shown in Figure 9.6. Consider a cylindrical volume element d"V of cross-sectional area
Figure 9.6. Geometry of the problem of scattering in a spherical atmosphere.
Center of Planet
9.E. Spherical geometry
257
where s' is the path of the incident ray. The slant-path optical depth for the emergent ray at that point is re(s) = fN[^"{s\s)\aQEds\ (9.35b) •'o where s" is the path of the emergent ray. Then the singly scattered radiance reaching the detector is
IDS
J ^ f f
= ^
^
(9.36)
Equation (9.36) is analogous to the Lommel-Seeliger expression (8.34) for plane atmospheres. From Figure 9.6,
& = [Ml +2s<^0cos e0 + s2f/2 s2 = &0 + s cos e0 + -Tygr sin2e0 + • • •.
(9.37)
Now, the distance s over which the integrand is appreciable is of the order of a few times %?. Hence, if cos e0 is not small, only thefirsttwo terms on the right-hand side of (9.37) need be considered. However, near the limb the second term can be zero, so that the third term must be included in the calculation of rt and IDs. Because higher terms are of the order of higher powers of ^ / ^ 0 , they may be ignored if this quantity is small. More precisely (Wallach and Hapke, 1985), higherorder terms may be neglected if ( 2 ^ / ^ 0 ) 1 / 2
s"2 s i n 2 " - & + s"cos e + YW~ «o' , 2 .2 cos e ~ cos e0 + 7™- 'sin e0, S
cos I ~ cos l0 + ^=r cos g.
(9.38) ( 9 - 39 ) (9.40) (9.41)
With these approximations the integrations in (9.35) can be carried
258
9. The bidirectional reflectance in other geometries
out to give ,1/2 TAS)
=
sin e0 Xexp -
Xexp
cos e sin en (9.42)
- + TTS? cot2 e0 erfcl
COS*
-•
sin
(9.43) T.I — <
sin e0 \ (9.44)
where erfc(jc) is the complementary error function of argument x, erfc(jc) = 1 — erf(x), and erf(x) is the error function. Now, the density of scatterers N decreases exponentially with altitude, while the transmission of the atmosphere exp( -rt- re) increases exponentially with an exponential function of altitude. Hence, the integrand in (9.36) is sharply localized along the line of sight. Wallach and Hapke (1985) showed that the center of this active zone is located at a point s = s* given by the set of transcendental equations
. ^ .* cosg 2 + |cos e* -cos IT sin . 2 i0\ sinll0 Xexp
cos 2 6*
sin
erfc
Sin
cos er (9.45a)
9.E. Spherical geometry
259
where cos e* = cos e0 + -^- sin2e0,
(9.45b)
cos 6 = cos i0 + j ~ - cos g.
(9.45c)
Equations (9.45) may be rapidly solved for 5* by setting s* = 0 in the argument of the logarithm in (9.45a) and iterating. A useful approximation for the error function is
erf(±*)«±{l
_
r
J~X
_,
r^}.
(9.46)
This approximation is accurate to better than 1% everywhere. Wallach and Hapke also noted that the ratio T^s)/rt{s) is very nearly constant throughout the active zone, so that (9.36) may be evaluated approximately by setting it equal to its value at 5 = 5*. Then,
w
Over most of the disk of the planet, where neither cos LQ nor cos e0 is small, equation (9.47) becomes
which is the usual Lommel-Seeliger expression. Near the limb, where cos e0 is small, (9.47) has the asymptotic behavior /D,«/^P(*){1-«P[-T
#
(-OO)]},
(9.49)
which, according to (9.44), decreases with altitude proportional to exp[ — (^Q — 31^)/&\ In this region a semilog plot of brightness with altitude is a straight line with slope — ^ ~ x , and the intensity is proportional to the integral of N{^)aQsp{g) along the line of sight.
260 9. The bidirectional reflectance in other geometries Near the terminator, where cos Co is small, (9.47) becomes approximately
P(g)eM' [&(**)- ^ j / ^ J s e c e0,
(9.50)
so that the intensity is proportional to the density of scatterers at the center of the active zone. 9.E.2. Effects of multiple scattering Wallach and Hapke (1985) also derived an approximate expression for the contribution of multiple scattering, and that reference should be consulted for details of the derivation. However, for small regions near the limb or terminator, where the curvature of the planet must be taken into account, the main effect of multiple scattering is to multiply IDs by a constant term. An example of these equations fitted to spacecraft observations of the terminator of Venus is shown in Figure 9.7. The scale height of the aerosols in that region of the atmosphere was found to be ^ = 9.2 km. Figure 9.7. Relative brightness in the vicinity of the terminator on Venus. The heavy stepped line shows the radiance measured by the Mariner 10 spacecraft. The straight lines show the theoretical brightness for two different aerosol scale heights %? calculated from equation (9.50). From Wallach and Hapke (1985), courtesy of Academic Press, Inc.
40 80 120 DISTANCE ON REFERENCE SURFACE, km
10 Other quantities related to reflectance, integrated reflectances, planetary photometry, reflectances of mixtures 10.A. Introduction
In Chapter 8, expressions for the bidirectional reflectance of a particulate medium were derived, and variants in different geometries were discussed in Chapter 9. In this chapter we continue the treatment of reflectances by introducing several quantities that are frequently encountered when measuring reflectance and by deriving expressions for reflectances that involve integration over one or both hemispheres. The remote sensing of bodies of the solar system has its own nomenclature because of the special problems of astronomical observation. Expressions for quantities commonly encountered in planetary spectrophotometry will be derived. Finally, we will give formulas with which one can calculate the reflectances of media consisting of mixtures of different kinds of particles. 10.B. Bidirectional reflectance quantities The bidirectional-reflectance distribution function (BRDF) is the ratio
of the radiance scattered by a surface into a given direction to the collimated power incident on a unit area of the surface. The incident radiant power per unit area of surface is Jfi0, and the scattered radiance is Jr(l,e,g), where r(i, e,g) is the bidirectional reflectance 261
262
10. Other quantities related to reflectance
of the surface. Thus, BRDF(i,e,s)-r(i,e,*)//t 0 .
(10.1)
If r is described by (8.89),
(10.2) The reflectance factor, denoted by REFF, of a surface is defined as the ratio of the reflectance of the surface to that of a perfectly diffuse surface under the same conditions of illumination and measurement. The reflectance factor is sometimes also called the reflectance coefficient Because the bidirectional reflectance of a perfect Lambert surface is rL = fjiQ/TT, the reflectance factor of a surface with bidirectional reflectance r(i, e,g) is (10.3) If r is given by (8.89), then
(10.4) The radiance factor, denoted by RADF, is similar to the reflectance factor, except that it is defined as the ratio of the bidirectional reflectance of a surface to that of a perfectly diffuse surface illuminated at £ = 0, rather than at the same angle of illumination as the sample. The radiance factor is given by RAD F (I, t, g) =
IT r (i,
e, g)
(10.5) The relative reflectance, denoted by T, of a particulate sample is defined as the reflectance relative to that of standard surface consisting of an infinitely thick particulate medium of nonabsorbing, isotropic scatterers, with negligible opposition effect, and illuminated and viewed at the same geometry as the sample.
10.C. Reciprocity 263 If the bidirectional reflectances of the sample and the standard are given by (8.89), the relative bidirectional reflectance is
Using approximation (8.55) for the H function, the relative bidirectional reflectance of a sample of isotropic scatterers measured outside of the opposition peak is given by
Strictly speaking, the REFF, RADF, and T should be modified by the same adjectives as the reflectance to denote the illumination and viewing geometries. However, it follows directly from the law of conservation of energy that the directional-hemispherical and bihemispherical reflectances of any perfectly nonabsorbing surface are unity. It will be left as an exercise for the reader to show that the hemispherical-directional reflectances of a perfect Lambert surface and of a medium of nonabsorbing isotropic scatterers are also unity. Hence, the directional-hemispherical, hemispherical-directional, and bihemispherical reflectance factors, radiance factors, and relative reflectances of a surface are all equal to their respective reflectances. When the REFF, RADF, and T are unmodified by any prefixes, they have the general connotation of referring to the bidirectional reflectance. 10.C. Reciprocity
A powerful and useful theorem in reflectance work is the principle of reciprocity, which was first formulated by Helmholtz (Minnaert, 1941). The principle may be stated as follows: Suppose a uniform surface is illuminated by light from a collimated source making an angle I with the surface normal, and observed by a detector that measures the light emerging from a small part of the illuminated surface at an angle e from the normal. Let the bidirectional reflectance of the surface under these conditions be r(l, e,g). Next, interchange the positions of the source and detector, and denote the new reflectance by r(e,L,g). The reciprocity principle states that the reflectances must satisfy the following relation: , = r(e,l,g)cosL (10.8)
264
10. Other quantities related to reflectance
The proof is intuitive. Note that /rU, e,g)cos e is the radiance scattered by unit area of the surface when the source is at angle I and the detector at angle e with the normal. Similarly, Jr(e, &,g)cos I is the radiance scattered by a unit area in the reciprocal geometry. Imagine that the source and detector are connected by ray bundles of light, which emerge from the source, pass through this unit area of surface, are scattered within the medium, and pass into the detector. Although the rays follow a multitude of complex paths within the medium, these paths remain the same when the positions of the source and detector are interchanged. The property of reciprocity follows from the fact that photons can travel in either direction along the ray paths. Equivalent statements of the principle of reciprocity are ,;,g),
(10.9a)
REFF(;,e,g) = REFF(e,;,g).
(10.9b)
or The principle of reciprocity is useful for testing proposed scattering laws. If they do not obey reciprocity, they are incorrect. Note that (8.89) is reciprocal, as are the Lambert and Minnaert laws. Equation (10.8) may also be used to calculate the brightness of a surface at a given set of angles from measurements made at the reciprocal angles. One caveat must be emphasized: An important assumption underlying the principle is that the surface is laterally uniform. If the measured brightness of a surface is found to actually violate reciprocity, the likely cause is a lateral inhomogeneity, which causes the detector to view two different kinds of surfaces when source and detector are interchanged. 10.D. Integrated reflectances 10.D.1. Biconical reflectances
The basic expression for the bidirectional reflectance of a semiinfinite particulate medium was derived in Chapter 8 and is given by equation (8.89). If the source and detector do not occupy negligibly small solid angles as seen from the surface, appropriate expressions for the reflectances may be found by integrating (8.89) over the angular
10.D. Integrated reflectances
265
distribution of the radiance from the source and the angular distribution of the response of the detector. In general, such reflectances will be biconical. However, because they would be specific to each particular system, it would not be particularly useful to discuss biconical reflectances further here.
10.D.2. The hemispherical reflectance (directional-hemispherical reflectance)
The directional-hemispherical reflectance, or, more simply, the hemispherical reflectance, is denoted by rh. It is defined as the ratio of the total power scattered into the upper hemisphere by a unit area of the surface of the medium to the collimated power incident on the unit surface area. The hemispherical reflectance is also called the hemispherical albedo or the plane albedo in planetary photometric work, where it is denoted by Ah (see Section 10.E.4). The hemispherical reflectance is important for two reasons. First, it is the quantity that is measured by many commercial reflectance spectrometers. Second, it is one of the properties of a material that determines the radiative equilibrium temperature. The power incident on a unit area of the surface of a medium is JfiQ. The power emitted into unit solid angle per unit area of the surface is Y= Jr(l, e,g)fi. The total power emitted per unit area of the surface into the entire hemisphere above the surface is obtained by integrating the emitted power per unit area over the entire hemisphere into which the radiance is emitted. Hence, the general expression for the hemispherical reflectance is (
)
e
(10.10)
2TT
where dCle = sin e d e diff.
We wish to substitute (8.89) for r in this integral. However, before proceeding, note that the opposition effect typically has an angular half-width of about 0.1 radian or less. Hence, its relative contribution to the integral over a hemisphere is < 7r(0.1) 250/27r «: 1, so that the effect makes a negligible contribution to the integrated reflectance. Thus, to a sufficient approximation, B(g) may be ignored in the integrand.
266
10. Other quantities related to reflectance
Substituting equation (8.89) (with Bo = 0) for r in (10.10) gives
(10.11) It is convenient to separate the integral into two parts. Let where \A)
/»1
II
(10.12)
is the contribution to rh by an equivalent medium of isotropic scatterers, and
is the part of rh due to anisotropy. The isotropic component is readily evaluated by writing ja/(/x 0 + n) = 1 - M O /(/X O + jit). Then, using (8.50) and (8.53),
= l-yif(/t0).
(10.14)
Equation (10.13) will be evaluated for the case where p(g) can be expressed as a first-order Legendre polynomial: p(g) = l + b1cosg = l+b^ofi
+
(l-ti)1/\l-fi2)^
The integral over cos i// is zero. The remaining two terms are readily integrated to give
^ ]
(10.15)
Combining rhi and rha gives
[|
^ ] - (10-16)
Using approximation (8.55) for H(JJL0) and putting ju,oln[(ju,o + 1)/JLL0] — 2JJLO/(1 + 2/XO), which has the same accuracy as (8.55), (10.16)
267
10.D. Integrated reflectances
becomes (10.17) In Figure 10.1, rh(l) is plotted versus w for several values of I for the case of isotropic scatterers. An alternate derivation of (10.17) for the case of isotropic scatterers can be found from equations (8.43)-(8.45). In the two-stream approximation, the upward-going flux at the surface is 7,(0)=
Hence, the total power per unit area leaving the surface in the upward direction is TT/2
#
J f i l l p r \ c p / /rr c i x\ P ft P -—*
1V 7
1— Jii
y
0
^ l + 2yja0'
from which it follows that rh(l) = (l-y)/(l + 2yfjL0). This expression was first obtained by Reichman (1973). The directional-hemispherical reflectance of a medium of nonabsorbing particles must equal unity. However, equation (10.17) predicts Figure 10.1. Hemispherical reflectance rh (also called directionalhemispherical reflectance, hemispherical albedo, and plane albedo) for isotropic scatterers plotted against the single-scattering albedo for several values of the angle of incidence. The curve for I = 60° is identical with r 0 . The curves of the hemispherical-directional reflectance are identical if I is replaced by e.
0.8
0.6
0.4
0.2
0.8
268
10. Other quantities related to reflectance
that when w = 1, rh{ l) = 14- b1/ji0 /4(1 + 2JLI0); for instance, when I = 0, rh = l + bt /12, so that if bx is as large as 1, the discrepancy is 8%. This is an indication of the magnitude of the general errors inherent in the approximation that places all the effects of particle anisotropy in the single-scattering term of the bidirectional reflectance.
10.D.3. The hemispherical-directional reflectance
The hemispherical-directional reflectance is defined as the ratio of (1) the radiance scattered into a particular direction from the surface of a medium that is being uniformly illuminated from all directions in the hemisphere above the surface to (2) the incident radiance. To calculate the hemispherical-directional reflectance, the incident irradiance / must be replaced by / o rfn., where 70 is the incident radiance, assumed independent of direction, and dft. = sin idldijj. The radiance scattered from the surface into a given direction is the integral of Ior(i,e,g) dil. over the hemisphere from which the incident radiance comes. Hence, the hemispherical-directional reflectance is
r l
o ( )
i
f
( g ) ,
(10.18)
By the same arguments as used in deriving the hemispherical reflectance, the contribution of the opposition effect to the total scattered radiance is negligible. Inserting (8.89) with J50 = 0 into (10.18) gives an expression that is identical with (10.11), except that I and e are interchanged. Hence, the hemispherical-directional reflectance has the same functional dependence on e that the hemispherical reflectance has on i, and we may write, by inspection, (10.19) Making the same approximations as for the hemispherical reflectance gives the approximate form
i4
<1020)
-
If I is replaced by e in Figure 10.1, the curve shows rhd as a function of w for several values of e.
10.D. Integrated reflectances
269
10.D.4. The spherical reflectance (bihemispherical reflectance)
The bihemispherical reflectance, or, more simply, the spherical reflectance, is denoted by rs. It is the ratio of (1) the total power scattered into the upward hemisphere from unit area of a surface that is being uniformly illuminated by diffuse radiance from the entire upper hemisphere to (2) the total power incident on unit area of the surface. In Section 10.E.7 this quantity will be shown to be equivalent to the Bond or spherical albedo As of a spherical planet. The total power incident on unit area of the surface is /2/7r/0ju,027rsin i di = IOTT, where / 0 is the incident radiance. The total power scattered into the upper hemisphere is the integral of IQ dCl. r/jb dile over both the direction from which the radiance comes and the direction into which it is scattered. Hence, the spherical reflectance is
Equation (10.21) will be evaluated first for the case of isotropic scatterers. Then r is independent of «/f, and we may put dd& = 2TT d/x and dft< = 2TT dfi0. Using (8.89) with p(g) = 1 and Bo = 0, 1
r5 = — /
r l /»1 W
II
/ -r
UnM
r^Kjr^
TT/
——H(iA
TT J r\ J r\ ^77* jLtn
i fJL
\ TT/
\^
0)Ji(iJL)27r
JO
J
ay^lir a/i 0
where we have used the result from equation (10.14). Hence, r, = l-2yHl9
(10.22)
where Ht is the first moment of the H function {equation (8.52)]. The quantities Hx and rs have been calculated by numerical integration by Chamberlain and Smith (1970). The spherical reflectance is plotted against w in Figure 10.2. An approximate expression for rs may be obtained by using (8.58b) for Hx in (10.22) to obtain \
\
^
\
(10.23)
This approximation has relative errors of less than 2.0%. Note that rs and r0 are both types of bihemispherical reflectances and have very similar functional dependences on w; however, they are
270
10. Other quantities related to reflectance
not equal. The reason has to do with the way the incident radiance is assumed to interact with the medium in the derivations of the two expressions. In the diffusive reflectance r0 the incidence radiance is a boundary condition that interacts only via the imaginary "membrane" at the mathematical upper surface. However, in the more physically realistic solution for rs, the interaction takes place via the source function within a layer a few times \/E thick below the upper surface. This alters the distribution with depth of the radiance within the medium, and hence alters the reflectance. However, it was emphasized in Chapter 8 that the diffusive reflectance is a useful, mathematically simple quantity that gives surprisingly accurate first-order estimates of reflectance. As a demonstration of the power of the diffusive reflectance, it is left as an exercise for the reader to show that the following identities hold for media of isotropic scatterers: (1)
r ( 6 0 ° , 6 0 ° , g » / 0 = 0r,
(10.24a)
(2)
^(60°) = r0 ,
(10.24b)
(3)
^ ( 6 0 ° ) = r0 .
(10.24c)
Because of the similar behaviors of the diffusive and spherical reflectances, it is reasonable to ask if similarity relations can be found Figure 10.2. Spherical reflectance rs (also called bihemispherical reflectance and Bond albedo) versus w for isotropic scatterers. The exact and approximate expressions are indistinguishable on this scale. Also shown is the diffusive reflectance r0.
0.8
10.E. Planetary photometry
111
that will convert the expression for rs for isotropic scatterers to one applicable to nonisotropic scatterers. Although a general answer cannot be given, this question has been investigated numerically by Van de Hulst (1974). He calculated the spherical reflectance of a medium of scatterers having an angular function that can be described by the Henyey-Greenstein function, V
'
(l
where £ is the cosine asymmetry factor. Van de Hulst finds that replacing w and y by w* and y*, respectively, in (10.22), where these quantities are given by the similarity relations "* = jzjZW,
(10.25a) (10 25b)
gives an approximation for rs that is accurate to 0.002 for all values of w and £. 10.E. Planetary photometry 10.E.1. Introduction
Astrophysicists usually describe the photometric properties of solarsystem bodies by several different kinds of reflectances known as albedos and photometric functions. The word albedo comes from the Latin word for whiteness. Just as there are many reflectances, so there are several different kinds of albedos, depending on the geometry. Historically, most of these quantities arose in attempts to characterize the scattering properties of the surface of the moon as observed from the earth, and the definitions were then extended to other bodies of the solar system. In addition to defining the various quantities for a planet whose surface has arbitrary photometric properties, the following sections will derive approximate analytic expressions for these quantities on a spherical body covered with an optically thick, uniform, particulate regolith whose bidirectional reflectance can be described by an equation of the form of (8.89).
272
10. Other quantities related to reflectance
As usual, rU, e,g) will denote the bidirectional reflectance of a surface, and / the incident collimated irradiance. The radiance I(i,e,g) interacting with the eye produces the sensation of brightness, and the terms radiance and brightness are often used interchangeably in planetary work. 10.E.2. The normal albedo
The normal albedo An is the ratio of the brightness of a surface observed at zero phase angle to the brightness of a perfectly diffuse surface located at the same position, but illuminated and observed perpendicularly. That is, the normal albedo is the radiance factor of the surface at zero phase. Thus, AH = [Jr(e,690)]/[J/ir]=irr(e,6,0).
(10.26)
Using (8.89),
% [(1 + B0)p(0) + H(tf -1].
(10.27)
In general, the normal albedo is a function of e. However, for dark bodies of the solar system, including the moon, Mercury, and many asteroids, [H(/JL)2 - 1 ] «*: 1, so that AH~%(l
+ BQ)p(p),
(10.28)
which is constant, independent of e. Thus, the albedos of different areas on the surfaces of these bodies may be intercompared even though they are observed at different angles. Hence, the normal albedo is a useful parameter to characterize their surfaces. 10.E.3. The photometric Junction
The ratio of the brightness of a surface viewed at a fixed e, but varying I and g, to its value at g = 0 is called the photometric function of the surface: f(i,*,g) = r(i9e,g)/r(e,e90).
(10.29)
Then the radiance scattered by the surface is given by e,g) = -Anf(L,e,g). (10.30)
10.E, Planetary photometry
273
For a Lambert surface, An = AL, and f( i,, e, g) = JLI0. For a particulate medium that obeys (8.89),
)]rti)+«(*)g(iOi Because the moon keeps the same face toward the earth, a given area will always be viewed from the earth at nearly the same e. Because the lunar surface has a low albedo, H(IJLO)H(IJL) — 1 <: 1, and 2 - 1
This expression has the same functional dependence on i, e, and g as does the generalized Lommel-Seeliger law, equation (8.35a). KenKnight, Rosenberg, and Wehner (1967) have shown that the relative brightnesses of different areas on the moon are described to a good approximation by the Lommel-Seeliger law. 10.E.4. The hemispherical albedo (plane albedo)
The hemispherical albedo Ah\s the total fraction of collimated irradiance incident on unit area of surface that is scattered into the upward hemisphere. It is also known as the plane albedo. The hemispherical albedo is the same as the directional-hemispherical or hemispherical reflectance; that is, Ah{i) = rh{i).
(10.33)
Explicit expressions for this quantity have been derived, and analytic approximations are given in Section 10.D.2. It is plotted versus w and I in Figure 10.1. 10.E.5. The physical albedo (geometric albedo)
The physical albedo is also known as the geometric albedo. It is defined as the ratio of the brightness of a body at g = 0 to the brightness of a perfect Lambert disk of the same radius and at the same distance as the body, but illuminated and observed perpendicularly. The physical albedo is the weighted average of the normal albedo over the illuminated area of the body.
274
10. Other quantities related to reflectance
For a spherical body of radius & the physical albedo is given by
Jr(e,e,0)tidA\/[(J/7r)7r<%2]
A(00
J
(10.34)
[ AQ)
where dA = lir^sin ede= -2ir^2dfi is the increment of area on the surface of the body, and A(i) is the area of the illuminated hemisphere. Using (8.89),
= jf\[l
+ Bo}p(O)-l(10.35)
where Api is the physical albedo of a sphere covered by a medium of isotropic scatterers: (10.36) The physical albedo for isotropic scatterers is plotted in Figure 10.3. An approximate expression for Api may be found be using the linear approximation for the H function, equation (8.56). If terms of order Figure 10.3. Physical albedo Ap (also called geometric albedo) of a planet covered with a medium of isotropic scatterers versus w. The exact and approximate expressions are indistinguishable on this scale.
0.8
0.6
"
0.4
"
0.2
"
0.2
0.4
0.6
w
0.8
10.E. Planetary photometry A-Q/24
275
are neglected, the expression A ~ —r
4- — r
2
MO VT\
is obtained. This has relative errors of less than 3% everywhere. Because including the terms of order r^/24 does not decrease the errors, the simpler expression (10.37) will be used. Combining (10.35) and (10.37) gives
AD « r o (4 + i r 0 ) + £ [(1 + 50)p(0) -1].
(10.38)
Equation (10.37) may be empirically modified to give an expression for Api that is accurate to better than 0.4% everywhere: Api«
0.49r0 + 0.196r2.
(10.39)
Note that Ap may exceed unity if the opposition effect is large or if the particles of the medium are strongly backscattering. This simply means that the bidirectional reflectance of the medium is more backscattering than that of a Lambert surface and does not violate any restrictions imposed by conservation of energy. 10.E.6. The integral phase Junction
The integral phase function of a body is the relative brightness of the entire planet seen at a particular phase angle, normalized to its brightness at zero phase angle. Thus,
(
( ) | / [ [
( ) A
(10.40)
where A(i,v) is the area of the planet that is both illuminated and visible at phase angle g, and A(i) is the area of the illuminated part. If the body is a sphere of radius «£?, then, from (10.34), ^
/
(10.41)
We will derive an approximate expression for a spherical planet covered by a soil whose bidirectional reflectance can be described by (8.89). The phase function can be written in the form p,
(10.42)
276
10. Other quantities related to reflectance
where 4>t(g) is the isotropic portion of the phase function, (10 43)
-
and ®n(g) is the nonisotropic contribution, . (10.44) A quantity proportional to the last integral has already been evaluated in Chapter 6 for a sphere whose surface obeys the LommelSeeliger law. Using that result gives
(10.45) An approximate expression for <&,(#) may be obtained by using the linear approximation for the H functions, equation (8.56), and ignoring terms of order r^/24, as was done in the derivation of the approximation for Ap. The integration is straightforward and gives
*,(g)-^r0{[l-r0][l-sm§tan§ln(cot§)] Combining (10.42)-(10.46) gives
X [l - sin | tan | lnfcot |
where Ap is given by (10.38). The first term of (10.47) describes a sphere covered by a Lommel-Seeliger scattering surface, modified by the particle phase function and the opposition effect. This term will be dominant for low-albedo bodies, such as the moon. The second term describes a sphere covered by a Lambert scattering surface. This term will dominate for high-albedo bodies, such as Venus or icy satellites. Equations based on (10.47) have been used to describe the integral phase functions of a wide variety of bodies in the solar system (Hapke, 1984; Buratti, 1985; Helfenstein and Veverka, 1987; Simonelli and
10.E. Planetary photometry
277
Veverka, 1987; Kerbst, Skrutskie, and Nicholson, 1987; Veverka et al., 1988; Bowell et al., 1989; Domingue et al., 1991). In order to model light scattering from planetary bodies more realistically, equation (10.47) must be modified to include the shadowing and other phenomena caused by large-scale roughness. These modifications will be described in Chapter 12. 10.E.7. The spherical albedo (Bond albedo)
The spherical albedo, also called the Bond albedo, is the total fraction of incident irradiance scattered by a body into all directions. It bears the same relation to a planet as the single-scattering albedo w does to a particle. The spherical albedo may be found by integrating the fraction of power incident on unit area of a planet emitted into all directions Pdh(i) over the illuminated area A(i) of the body. For a sphere,
Jr(L,e,g)jjLd(ledA
(10.48)
Because the increment of the illuminated half of a sphere is dA = «$? 2sin
hi / r(i,e,g)ndncdCle.
(10.49)
This is identical with equation (10.21) for the bihemispherical or spherical reflectance rs. Thus, A , = ra9 (10.50) as stated without proof in Section 10.D.4. In particular, the spherical albedo of a medium of isotropic scatterers is As = l-2yHl (10.51a) and may be approximated by
^ H ^ )
d0.51b)
As is plotted as a function of w in Figure 10.2. For nonisotropic scatterers the similarity relations (10.25) may be used. Conservation of energy requires that the spherical albedo cannot exceed unity.
278
10. Other quantities related to reflectance 10.E.8. The bolometric albedo
The bolometric albedo Ab is the average of the spectral spherical albedo ^4/A) weighted by the spectral irradiance of the sun JS(X):
Ab = \fmAa(X)Jt(X) dx\ /\fjs(X)
dx].
(10.52)
l/o J/ l/o J /S(A) is approximately the Planck function of a black body at a temperature of 5,770°K. 10.E.9. The phase integral The phase integral q of a body is defined as $(g)singdg,
(10.53)
where d£lg = 2TT sin gdg is the increment of solid angle associated with the phase angle. An important relation involving q, As, and Ap may be derived as follows: From (10.41),
q = \[ f J
J
\_ 4ir A(i,v)
Jr(L,e,g)vLdAdng}
\j
[JTT<%2AP\.
Now, the integral in this last expression for q is the total amount of light scattered into all directions by the entire surface of the body. But, by the definition of the spherical albedo, this is just Jir9i2As. Hence, q = As/Ap. (10.54) Russell (1916) has noted the following empirical relation, which is known as Russell's rule:
10.F. Mixing formulas 10.F.1. Areal mixtures Few materials encountered in nature consist of only one type of particle. Hence, it is necessary to be able to calculate the reflectances of mixture of different particle types. In remote sensing, two kinds of mixtures are of interest: areal and intimate. Areal mixtures are also
10.F. Mixing formulas
279
known as inhomogeneous, linear, macroscopic, and checkerboard mix-
tures. In an areal mixture, the surface area viewed by the detector consists of several unresolved, smaller patches, each of which consists of a pure material. In this case the total reflectance is simply the linear sum of each reflectance weighted by area. That is, 10 55
>- = V y >
( - )
Here r may represent any type of reflectance, as appropriate, r. is the same type of reflectance for the /th area, and F. is the fraction of the area viewed by the detector that is occupied by the /th area. 10.F.2. Intimate mixtures
Intimate mixtures are also called homogeneous and microscopic mixtures. In an intimate mixture the surface consists of different types of particles mixed homogeneously together in close proximity. In this case the averaging process is on the level of the individual particle, and the parameters that appear in the equation of radiative transfer are averages of the properties of the various types of particles in the mixture weighted by cross-sectional area. Because the parameters enter into the reflectance equation nonlinearly, the reflectance of an intimate mixture is a nonlinear function of the reflectances of the pure end members, as has been noted empirically by several authors (e.g., Nash and Conel, 1974). The formulas for intimate mixtures has another important use: calculating the effects on the reflectance of small asperities and subsurface fractures at the surfaces of grains. It was shown in Chapter 6 that such structures may be treated as small particles mixed with the large ones. If the surface structures are smaller than the wavelength, they can act as Rayleigh scatterers and absorbers. If the imaginary part of the refractive index is small, they act primarily as scatterers and increase the reflectance. However, if k is large, as for a metal, the asperities are efficient Rayleigh absorbers and decrease the reflectance. The mixing formula for each parameter follows from the definition of that parameter, as given in Chapter 7. Because these formulas follow from the definitions in the radiative-transfer equation, they are independent of the method or type of approximation used to solve that equation.
280
10. Other quantities related to reflectance
Mixtures are often specified by the fractional mass of each component, rather than by the number of particles. Assume that the particles are equant, and let the subscript / refer to any property of the particles, such as size, shape or composition. Then the cross-sectional area of the / t h type of particle is (r+ = ira2p
(10.56)
where a^ is the radius of the / t h type of particle. The bulk density of the / t h component is jrf-Nf±ira*Pf,
(10.57)
where N. is the number of particles of type / per unit volume, and p. is their solid density. Then
where D^ = 2a^ is the equivalent size of the / t h type of particle. From equations (7.36)-(7.47), the average single-scattering albedo is
and the average particle scattering function is
-
Because WJ = QSJ/QEJ> the alternate forms
G(g)
S ~
X}N} 1,
the last two expressions may be written in
(10.62)
10.F. Mixing formulas
281
If there is a continuous distribution of properties, the summations in these definitions may be replaced by integrations. If (10.62) is multiplied by (l/47r)cosg and integrated over solid angle, the following equations for the average cosine asymmetry factor is obtained:
where ^ is the cosine asymmetry factor of the / t h type of particle. From equations (8.82),(8.86), and (8.87), the opposition-effect parameters are 5(0)
(10 64)
-
where 5^(0) is the amount of light scattered into zero phase from the back surface of the / t h particle, and
*-i (10.65) where the filling factor is
If the mixture is a binary one, a useful formula due to Helfenstein and Veverka (private communication, 1990) may be derived as follows: For a mixture of two types of particles, equations (10.61) and (10.63) may be written W
D?
i
TJf
anQ
C
D7T7I
I ~j ...
»
where J/issN1at1QE1 and Jf2 = N2a2QE2. These equations can be solved for s/x and sf2, and the resulting expressions substituted back
Figure 10.4. Comparison of the measured (solid lines) and calculated (dotted lines) spectra of 0.25/0.75, 0.5/0.5, and 0.75/0.25 intimate mixtures of olivine (OL) and hypersthene (HYP). The spectra are vertically offset to avoid confusion. From Johnson et al. (1983); copyright by the American Geophysical Union.
0.5
1.0
1.5
2.0
WAVELENGTH (^m)
Figure 10.5. Reflectances of intimate mixtures of montmorillonite and charcoal. The points are the measured values. The line shows the values calculated from mixing theory. From Clark (1983); copyright by the American Geophysical Union.
-3.0
-2.0
-1.0
log (weight fraction of carbon black)
0.0
10.F. Mixing formulas
283
into the expression for £, to give
Finally, the espat function of a mixture is
If the particles are large and closely packed, so that QE^ = 1, and
(10.69) where W^il-w^/w^ is the espat function of the /th type of particle. Note that the weighting parameter of W^ inW is Jf^Wj, /p^D^ not Jlk/pkDk. The mixing equations have been verified experimentally by several workers. Johnson et al. (1983) and Clark (1983) measured the reflectances of mixtures and compared them with the values predicted from the end members. Comparisons between their predicted and measured reflectances are given in Figures 10.4 and 10.5.
11 Reflectance spectroscopy
ll.A. Introduction
One of the objectives of studying a planet by reflectance is to infer certain properties of the surface by inverting the remote measurement. In the laboratory, the objective of a reflectance measurement is usually to determine the spectral absorption coefficient of the material or, at least, some quantity proportional to it, by inversion of the reflectance. There are at least three reasons why reflectance spectroscopy is a powerful technique for measuring the characteristic absorption spectrum of a particulate material. First, the dynamic range of the measurement is extremely large. Multiple scattering amplifies the contrast within very weak absorption bands in the light transmitted through the particles, while very strong bands can be detected by anomalous dispersion in radiation reflected from the particle surfaces. Hence, the measurement of a single spectrum can give information on the spectral absorption coefficient over a range of several orders of magnitude in a. Second, sample preparation is convenient and simply requires grinding the material to the desired degree of fineness and sieving it to constrain the particle size. Third, reflectance techniques are effective in the range k ^10" 3 -10~ 1 , where both transmission- and specular-reflection techniques are very difficult. By contrast, if a(A) is measured by transmission, the sample must be sliced into a thin section that must then be polished on both sides; also, the range by which a(A) can vary is limited to about one order of magnitude. 284
ll.B. Measurement of reflectances
285
The characteristics of a medium that can be measured directly by reflectance are those that occur in the expression for the bidirectional reflectance, equation (8.89), and are of two types: (1) those that describe the scattering properties of an average single particle, w, pig), and Bo; and (2) those that describe the physical structure of the surface. The latter category includes the porosity and particle size distribution, which determine the opposition-effect width parameter h and also the mean large-scale roughness angle 0, which will be derived in Chapter 12. In this chapter we will discuss the properties of a surface that can be studied using reflectance and the methods for inverting the reflectance data to estimate these properties. It will be seen that the single-scattering albedo w plays the central role. The reflectance model developed in the previous chapters of this book will be applied to the interpretation of spectral reflectance data. First, observational methods of measuring the various types of reflectances defined in Chapter 8 and derived in Chapters 8-10 will be described briefly. Next, we will see how to manipulate the reflectance data to retrieve the parameters that appear in the theory, especially the single-scattering albedo and spectral absorption coefficient. The relation between the imaginary part of the refractive index and the reflectance will be discussed in detail, including the shapes of absorption bands seen in reflectance. The effects of layers on band contrast will be considered. Finally, the widely used Kubelka-Munk model will be examined critically.
ll.B. Measurement of reflectances
In this section we shall briefly describe the methods by which the various types of reflectances and albedos described in Chapters 8-10 can be measured. For more details, one of the standard texts on reflectance spectroscopy, such as Wendtland and Hecht (1966) or Kortum (1969), should be consulted. To measure reflectance, both the scattered radiance and the incident intensity must be measured. This can be done by measuring the radiance scattered from the surface and then rotating the detector to measure the incident irradiance. However, under many experimental conditions this is inconvenient. Hence, the usual procedure is to calibrate the incident light by measuring the intensity reflected from a
286
11. Reflectance spectroscopy
standard material of known scattering properties. Then the reflectance is given by
Desirable properties of a standard are that it be very highly reflecting and stable with time. Standards in wide use in the near-UV/visible/near-IR region of the spectrum include finely ground MgO (Middleton and Sanders, 1951), BaSO4 (Grum and Luckey, 1968), and polytetrafluoroethylene (PTFE) (Weidner and Hsia, 1981; Weidner, Hsia, and Adams, 1985; Fairchild and Daoust, 1988). The latter compound is available under the trade name Halon (Allied Chemical Co.) and, as of the date of this book, is the most widely used standard. Depending on the wavelength, materials used as standards in the IR include sulfur, gold, and KBr (Nash, 1986; Salisbury, Hapke and Eastes, 1987). The bidirectional reflectance can be measured by illuminating a material with light from a source of small angular aperture, as seen from the surface, and observing the scattered radiance with a detector that also subtends a small angle from the surface. To adequately constrain the scattering parameters, it is desirable that the reflectance be measured at many values of i, e, and g; however, frequently it is measured at only one fixed set of angles. The hemispherical reflectance {directional-hemispherical reflectance)
is measured using an integrating sphere. This device consists of a hollow cavity whose inner walls are covered with a highly reflective, diffusely scattering paint, perforated by two small openings, or ports, one for admitting the incident irradiance and another for observing the radiance inside the sphere. The material whose reflectance is being measured is placed inside the sphere, where it is illuminated by collimated light. The scattered radiance is sampled by a detector that views the inside of the sphere, but not the material. For accurate values, the measurements must be corrected for losses out of the ports (Hisdal, 1965; Kortum, 1969). The hemispherical-directional reflectance of a material can be mea-
sured using an integrating sphere operating in the reverse direction from that used to measure the directional-hemispherical reflectance. That is, light from the source does not illuminate the material directly, but only after multiple diffuse scatterings inside the sphere. The
ll.B. Measurement of reflectances
287
radiance scattered into a specific direction by the sample is viewed directly by a detector of small angular aperture. The bihemispherical reflectance or spherical reflectance of a material can, in principle, be measured by covering an opaque sphere (actually, a hemisphere is sufficient) with an optically thick coating of the sample and placing it at the center of an integrating sphere. One side of the sample is then illuminated by collimated light, and the radiance that is scattered in all directions is measured by a detector that does not view the target directly. It was shown in Section 10.E.7 that the bihemispherical reflectance of a flat surface illuminated by a uniform flux from an entire hemisphere is equivalent to the spherical albedo of a sphere illuminated by collimated irradiance from one direction. I am not aware of any published values of bihemispherical reflectances that have actually been measured in this way in the laboratory. However, Bond or spherical albedos are routinely determined for planets, satellites, and other bodies of the solar system. The diffusive reflectance should not be confused with the bihemispherical reflectance. The diffusive reflectance, which includes the Kubelka-Munk reflectance (see Section ll.H) as a special case, is strictly a mathematical artifice with no direct physical meaning and thus cannot be measured. The physical albedo or geometric albedo of a body of the solar system can, in principle, be determined by measuring its brightness at zero phase angle. Usually the intensity is calibrated by observing comparison stars of known brightness. Using the intensity of sunlight, which is also known, the brightness of a perfectly diffusing (Lambert) disk of the same radius and in the same position as the body can be calculated. The physical albedo is the ratio of the brightness of the body to that of the disk. Unfortunately, orbits and trajectories of both natural bodies and artificial spacecraft are such that few bodies are ever observed exactly at zero phase angle. Hence, the brightness of an object must be measured at other phase angles and extrapolated to zero phase. The opposition effect complicates the extrapolation, and several conventions for carrying it out are in use. One convention is based on the fact that if the brightness of a body as a function of phase angle is plotted in astronomical magnitudes, which is a logarithmic scale (see any textbook on astronomy), the data very nearly fall on a straight line over a limited range of phase angles. The opposition brightness is
288
11. Reflectance spectroscopy
determined by extrapolating to zero phase. Some workers ignore the opposition effect, whereas others include it by using either a linear logarithmic extrapolation of small-phase-angle data or a theoretical equation such as (8.89). When it is necessary to use a physical albedo, the convention by which it was calculated should be ascertained. The spherical albedo or Bond albedo is determined by measuring the integrated brightness of a body at several phase angles and numerically calculating the phase integral and physical albedo. The spherical albedo is then calculated using equation (10.54). The normal albedo of an area on the surface of a body is determined in the same way as the physical albedo. The brightness is measured at several phase angles to yield the photometric function of the surface, which is then extrapolated to zero phase angle, and the normal albedo is calculated. ll.C. Inverting the reflectance to find the scattering parameters ll.C.l.
Numerical inversion methods
If a large enough data set is available, the reflectance can, in principle, be inverted to find all of the parameters in the reflectance function. A major difficulty is that the parameters enter the function nonlinearly. Frequently a data set will consist of measurements of the bidirectional reflectance of a surface or the integral phase function of a planet at a much larger number of angles than the total number of parameters in the relevant expression for the reflectance. In this case the problem of finding the parameters is overdetermined. The usual criterion for finding the best fit to an overdetermined data set is to minimize the root-mean-square (rms) residual between the calculated and measured data points. Certain of the parameters are most strongly influenced by the angles at which the reflectance is small. Hence, in finding the best fit, it is important that points of low intensity be given the same weight as points of high intensity. This can be done by using either relative or logarithmic residuals, rather than linear residuals. That is, the rms residual erms may be defined either as 1/2
(11.2a)
11.C. Inverting the reflectance
289
or as jt
l/2
£ (log/ c/ -logi d >) /J?
I1/2
,
(ii.2b)
where /rf^ is the intensity of the /th data point, Ic. is the intensity calculated for that geometry using the appropriate equation, and Jt is the number of points. In planetary photometry the intensity is frequently given in astronomical magnitudes, which is a logarithmic scale, so that (11.2b) is an appropriate form. When the single-particle angular-scattering function p(g) is found by fitting equation (8.89) or one of the expressions derived from it to the data, an important caveat must be kept in mind. In equation (8.89) all of the effects of particle anisotropy have been placed in the single-scattering term also, whereas in a rigorous formulation the multiple-scattering term exhibits some of the effects. Hence, when a data set is inverted, the calculated p(g) overestimates the anisotropy (Figure 8.9). A few inversion methods will now be discussed. The trial-and-error method. This method can be used on a personal computer with a plotting program. The data points and the appropriate equation with trial parameters are entered into the plotting routine, and a trial curve is generated. Of all the parameters in the equation of radiative transfer, the reflectance is dominated by the single-scattering albedo H>(A) or, equivalently, the albedo factor y(A) (Hapke, 1981; Helfenstein and Veverka, 1989). This parameter determines both the amplitude of the reflectance and its general variation with angle. The opposition effect is important only at small phase angles, the large-scale-roughness parameter has the greatest effect at large phase angles, and p(g) essentially fine-tunes the angular dependence of the reflectance. Hence, with a little practice, the effect of changing each parameter becomes intuitive, and satisfactory fits can be found rapidly. This method was used by Hapke and Wells (1981), by Domingue and Hapke (1989), and by Domingue et al. (1991). Automated computer fitting routines. The rms residual can be re-
garded as a surface in a multidimensional space in which each parameter is a coordinate. Mathematically, the objective of finding the best fit is that of finding the lowest minimum on this surface. There are several techniques for minimizing the residuals in such types of problems, as discussed in standard textbooks (e.g., Bevington, 1969).
290
11. Reflectance spectroscopy
One method is to set up a grid in the parameter space and search all points on the grid. The degree of fineness of such a search is limited by the available computer time and memory. More sophisticated methods have been developed by Helfenstein (1986; Helfenstein and Veverka, 1987). Pinty et al. (1989) and Pinty and Verstraete (1991) have used the standard computer routine EO4JAF from the Numerical Algorithm Group (NAG) library. A problem with such methods is that they may find a local minimum in the residual, rather than the deepest global minimum. Hence, independent checks for uniqueness should be made whenever possible (Domingue and Hapke, 1989). Several examples of theoretical curves that were fitted to observational or experimental data were given in Chapters 8-10, and more examples will be given in this chapter and in Chapter 12. 11.C.2. Analytic solution for the single-scattering albedo
We have seen that w is the primary quantity that can be measured using reflectance. In this section, equations for finding w from a reflectance will be derived. In Section ll.D, methods for estimating the absorption coefficient a from w will be discussed using the model for single-particle scattering developed in Chapter 6. Frequently the reflectances of both the sample and standard are measured in only one geometric configuration. In that case, w is the only parameter that can be found, and in order to estimate it, some assumptions must be made concerning the other parameters. The simplest assumption that can be made concerning p(g) is that the particles scatter isotropically, so that p(g) = 1. If the measurement is of the bidirectional reflectance, the angles usually are outside the opposition peak, whereas if the measurement is of an integral reflectance, the opposition effect has little influence; hence, the opposition effect often can be ignored. (An exception to the last statement is when the measured quantity is the physical albedo.) Usually the intensities of light scattered from the sample and from a calibrated standard are measured, and the ratio of the brightness of the sample to that of the standard is calculated. Often it is assumed that the standard scatters like a perfect diffuse surface, so that this ratio is interpreted as the reflectance factor (REFF). However, there is no physical justification for this assumption. It is far more reasonable physically, as well as more internally consistent, to assume that
11. C. Inverting the reflectance
291
the standard scatters light like a particulate medium with H> = 1, pig) = 1, and B(g) = 0, so that the ratio should be interpreted as the relative reflectance T. Hence, it will be assumed in this chapter that the ratio of the reflectance of the sample to that of the standard is T. In this section the analytic approximations for the various types of relative reflectances developed in previous chapters will be solved for y, assuming that pig) = 1 and Big) = 0. The equations are easier to solve for y than for w. Once y is found, the single-scattering albedo can be calculated from w = l-y2. (11.3) The diffusive reflectance. We have already encountered one set of inversion formulas in equations (8.27) and (8.28) for calculating y from ro. Repeating these, for completeness, ( w
^-x. (11.4b) 2 (l + '-o) However, because r0 has no direct physical meaning, these equations are of limited value. Their main use is for making rapid semiquantitative estimates of y or w, rather than for accurate data reduction. The bidirectional reflectance. It will be assumed that the intensity is measured at a large enough phase angle that the opposition surge is negligible. Then the relative bidirectional reflectance is given by equation (10.7),
Solving this for y gives 7(1}
~ (11.6)
The hemispherical reflectance (directional-hemispherical reflectance).
The relative hemispherical reflectance of a medium of isotropically scattering particles is equal to the hemispherical reflectance, equation
292
11. Reflectance spectroscopy
(10.17),
which can be solved for y to give y(rh) = i The hemispherical-directional reflectance. From the symmetry between rhd and rh we may write, by inspection,
The spherical reflectance (bihemispherical reflectance). The relative spherical reflectance is equal to the spherical reflectance, equation (10.23),
which can be solved for y to give 2(3r,+2) ll.D. Estimating the absorption coefficient from the single-scattering albedo 11.D.1. Introduction
In order to find the absorption coefficient a, or, equivalently, the imaginary part of the index of refraction k, from reflectance, the single-scattering albedo w must first be found using one of the methods described in the preceding section. If the material is a mixture, its reflectance must be deconvolved to determine the singlescattering albedos of the components. Deconvolution is discussed in Section ll.F. Next, a suitable model, such as those described in Chapters 5-7, must be chosen to relate the single-scattering albedo to the fundamental properties of the scatterers, which are the refractive index m = n + ik and the particle size D and shape.
ll.D. Estimating the absorption coefficient
293
In laboratory studies it is often possible to determine, or at least estimate, all of the particle parameters except k, which may then be calculated. The size and shape can be estimated from microscopic examination of the particles, using either an optical or scanning electron microscope. The real part of the index of refraction may be available in handbooks if the identity of the material is known, or it can be determined by a number of standard methods, such as measurement of the Brewster angle, or the Becke-line method (Bloss, 1961). Even in remote-sensing measurements there may be clues, such as the distinctive wavelengths of the absorption bands, to some of these parameters. 11.D.2. Monominerallic paniculate media with X » / ll.D.Za. The relation between reflectance and absorption coefficient We will first consider a medium that is pure and consists of one type of particle, and for which the mean particles have X^>1. Mixtures will be discussed later. If only one type of particle is present and X*>1, then * = NEaQs/NEaQE = QS/QE = Qt9 (11.12) because QE = 1 and Qs = Qs for large particles in a medium in which the particles are in contact. Also, (11.13) According to the results of Chapter 6, when X » 1 the scattering efficiency of an equant particle is given by equation (6.20): (11.14) where Se and S, are the average Fresnel reflection coefficients for externally and internally incident light, respectively, and @ is the internal-transmission factor. It was shown in Chapter 6 that @ can be described by the expression
294
11. Reflectance spectroscopy
where 6 is the internal scattering coefficient inside the particle,
ri-[l-}/a/(a+
*)]/[l
+ j a / ( a + *)],
(11.15b)
and (D) is the effective particle thickness, defined as the average distance traveled by rays that traverse the particle once without being internally scattered. If there are no internal or near-surface scatterers, so that 6 ^ 0, then ® = e~«< D>.
(11.15c)
An important problem in the absolute measurement of a by reflectance is the mean path length (D) through the particle. In Chapter 5 it was shown that for a perfect sphere, (D) = 0.9£>. For a distribution of irregular particles, (D) is expected to be of the same magnitude as, but smaller than, the size. There are several reasons for this. First, as discussed in Chapter 6, almost any departure from sphericity decreases (D). Second, most natural media consist of a distribution of particle sizes. Reflectance measurements emphasize the brightest particles, which are the smaller particles if they are not opaque. Third, the mixing formulas weight the smallest particles the most. Hence, when (D) is determined by reflectance, it will usually correspond to the smallest particles in the distribution, rather than to sdme mean size. This was seen in Figures 6.17 and 6.20. The dependence of the reflectance on a and k will now be examined in detail. The diffusive reflectance r0 will be used in this discussion because of its mathematical simplicity and because its behavior is representative of all the types of reflectances. If desired, r 0 may be interpreted as r(60°,60°,g » h ) or as ^(60°) [equation (10.24)]. Figures 11.1 and 11.2 show the dependence of the single-scattering albedo w and the reflectance r0 on the absorption. To facilitate understanding, the curves are plotted in two ways: by using as the independent variable the imaginary part of the refractive index k (Figures 11.1a and 11.2a) and the effective absorption thickness a{D) (Figures 11.1b and 11.2b). In these figures it has been assumed that a{D) = 4kX. For purposes of illustration, the various parameters have been taken to have the following values: A = 0.5 jam, (D) = 50 jLtm (corresponding to ^ = 3 0 0 ) and 500 fim (corresponding to X = 3,000), 5 = 0.06 jim" 1 , and n = 1.50.
ll.D. Estimating the absorption coefficient
295
In general, three different reflectance regimes may be distinguished. These will be referred to as the volume scattering region, the weak surface-scattering region, and the strong surface-scattering region. The volume-scattering region. When a(D)
show that w and r0 are both close to 1. Note, however, that even for k ~ 10~7, which for the 500-/im particles corresponds to a(D) ~ 10~3, Figure 11.1. Single-scattering albedo w — Qs for particles of sizes and internal scattering coefficients indicated. The real part of the refractive index is n = 1.50. (Top) w versus k; (bottom) w versus a(D).
0.1
10"
10"
0.001 k
0.1
10
w
0.1
n = 1.50 ..i
0.001
. .......i
0.1
...i
.
10 <x
i
1000
'100000
296
11. Reflectance spectroscopy
there is an easily measurable difference between r0 and 1.0, so that the magnitude of the reflectance is sensitive to k even for small values of a. When a(D) < 3, the reflectance is dominated by light that has been refracted, transmitted, and scattered within the volume of the particle. Hence, this part of the curve is called the volume-scattering region. In the part of this region where <*(£>>< 0.1, w and r0 decrease slowly as k increases. The rate of decrease becomes more Figure 11.2. Diffusive reflectance r 0 for media composed of the particles in Figure 11.1. (Top) r0 versus k\ (bottom) r0 versus a(D).
0.1
:
0.01 10"
10"
0.001
0.1
10
k
0.1
s=0 s = .06 urn
n = 1.50
0.01 0.001
0.1
10
a
1000
100000
ll.D. Estimating the absorption coefficient
297
rapid when 0.1 < a(D) < 3, so that the reflectance is highly sensitive to k there. The reflectance is equally sensitive to particle size: As (D) increases, the part of the ro-versus-£ curve in the volume-scattering region shifts to the left, and the reflectance decreases. This dependence of reflectance on particle size has been noted by many authors (e.g., Adams and Felice, 1967). The weak surface-scattering region. When 4 = 0 and a(D)>3, the particles are essentially opaque, and all of the scattering occurs by reflection from the surfaces of the particles, so that w ~ Se, independently of the sizes of the particles. Now, Sea[(n-1)2 + k2]/ [(n + 1)2 + k2]. Hence, when k < 0.1, Se is determined entirely by n. The weak surface-scattering region occurs between the place where a(D) > 3 and k < 0.1. Here the curves of w and r0 are flat, so that k cannot be determined by reflectance, but can only be placed between upper and lower limits. However, adding surface asperities or internal scatterers, parameterized by 4, increases the reflectance and extends the volume-scattering region to larger values of k. The strong surface-scattering region. When k > 0.1, Se is now sensitive
to k, and both w and r0 increase with increasing k in this region. If X » 1 and the particle surface is relatively smooth, the reflectance is independent of particle size. However, if the surfaces of the large particles are covered with scratches, edges, or asperities that can act like Rayleigh absorbers in the strong surface-scattering region, the reflectance will be smaller than it would if the surfaces were smooth. Unusual scattering and absorption can occur if m2 ~ — 2, as discussed in Chapter 5. ll.D.Zb. Solving for a directly Solving (11.14) for © gives @
= e i-s e-si+siw
( u - 16 )
If 4 = 0, @ = e~a
298
11. Reflectance spectroscopy
then (D) can be estimated by assuming that it is equal to the smallest particles in the distribution. If n is known and the reflectance is in the volume-scattering region, then Se and S( can be calculated from n using the equations given in Chapter 6. Then a is given by (1U7)
If 6 is not small, then a and 5 can be found by measuring the reflectances of powders of the same material, but of different particle sizes. From the measured values of w for large and small particles, @ can be calculated. Then one has two different equations of the form of (11.15) for two unknowns, from which a and 6 can be found by iteration. Although the reflectance is sensitive to a over the entire volumescattering region, the part of the curve where a{D)
(11.18)
was defined in Chapter 6. It was shown that this quantity is approximately proportional to a over a larger range of a(D), so that in this linear region we may write W~aDe,
(11.19)
where De is an effective particle size of the order of twice the actual particle size. Combining (11.18) and (11.19) led to an approximate expression for w that is valid in the linear region, (1
ll.D. Estimating the absorption coefficient
299
W is plotted against a(D) in Figure 11.4, which shows the linear region and also the departures from linearity. The espat function has been used to analyze remotely sensed photometric data on Jupiter's satellite Europa by Johnson et al. (1988). Figure 11.3. Absorption efficiency QA versus a(D) for the particles in Figure 11.1. The straight line has unit slope. 10 s = 0
s = .06 jim"1 "
Q A o.i
0.01
n = 1.50 0.001 0.001
0.1
10
1000
10°
Figure 11.4. Espat function W versus a(D) for the particles in Figure 11.1. The straight line has unit slope. 100
10
w 0.1
0.01
0.001 0.001
0.1
10 a
1000
10°
300
11. Reflectance spectroscopy
If the medium consists of a mixture of particles (either by composition or size or both), then the espat function that will be deduced from the measured volume single-scattering albedo is, from equation (10.69),
where the subscript / denotes the property of the /th type of particle. For a monominerallic medium with a small particle size distribution the weighting functions in (11.21) are equal, and W = W -aDe. If a(A) is known in some spectral region, then De of a powder can be found by measuring W(X) of the powder at the same wavelengths, allowing the measurement of a(A) by reflectance over the entire range of wavelengths in which the particles are volume scatterers. Two important limitations of the espat function must be emphasized. First, Figure 11.4 shows that W is linearly proportional to a only when neither a{D) nor $(D) is large; that is, the particles must not be optically thick. Second, if a medium is not monominerallic or if it has a wide particle size distribution, then (11.21) shows that the volume average espat function is proportional to the weighted sum of the absorption coefficients of the individual components of a mixture only if aDe
ll.D. Estimating the absorption coefficient
301
optical path length as (11.22) If a a - l n r , then the MOPL should be a constant for a given material as a changes. Let us see if the MOPL is approximately constant anywhere in the volume-scattering region. If so, then — In r should be approximately proportional to a, which would be a useful relation. We shall again use the diffusive reflectance r0 because its behavior is representative of most types of reflectances, and it is mathematically simple. In the volume-scattering region, w =* 1/(1 + ctDe). If aDe
-2
"
-4
"
-6
"
-8
-
-10 2 a
plotted
302
11. Reflectance spectroscopy
MOPL is even roughly constant. Hence, neither the effective absorbance nor the MOPL seems to be a particularly useful quantity especially when contrasted with the espat function. 11.D.3. Paniculate media with X « : 1
The case of a medium of solid particles for which the mean X
t 250-500
10
—
/
i
'"id
/ (
/ I -
0
/
y
0-5 urn sifted
!
, -
0-5 urn packed
/ \
/
/
.
.
•
•
-
^ -
i
1
1
10
11
1" 12
14
ll.E. Absorption bands in reflectance 303 This discussion suggests that if k <^c 1, then the espat function W should be proportional to a, even for small particles. However, the value of the constant of proportionality De is not clear; it may be of the order of A. The situation is even more confusing when k is not small. Under certain conditions, w
The wavelength at which a solid-state absorption band occurs is the major diagnostic characteristic for determining composition by remote sensing. Similarly, the depths of the absorption bands in the spectra of minerals can provide information on relative abundance. However, because the process of diffuse reflectance is nonlinear, both the band center and shape seen in reflectance may differ from those measured
304
11. Reflectance spectroscopy
by transmission. In the past this has caused some workers to unjustly reject reflectance spectroscopy as being unreliable. It is important that anyone who desires to use reflectance as a spectroscopic tool be aware of these effects. The shape of an absorption band seen in reflectance in a medium of particles with X » 1 may be understood qualitatively by examining one of the curves of reflectance versus imaginary refractive index in Figure 11.1. In the wing of an isolated band, k « 1 and r 0 ^ 1. As the wavelength or frequency moves toward the band center, k increases, the corresponding point on the curve moves to the right, and r 0 decreases. The point moves along the curve until it reaches the center of the band, where k is maximum. As the wavelength moves away from the center toward the other wing, k decreases, and the point retraces its path back along the curve to very small k. The shape of a band, including the slope of the spectrum at any wavelength, is determined by the reflectance regime in which the center of the band occurs, which depends on both a and (D). This is illustrated in Figure 11.7 for absorption bands whose shapes have been calculated using the Lorentz model [equations (3.16)]. If a{D) is small enough that all points of the band fall in the volume scattering region, then the reflectance spectrum of the band is Figure 11.7a. Reflectance spectrum of a particulate material with an absorption band of Lorentzian shape that is sufficiently weak that it is entirely within the volume-scattering region. The center of the band is located at vn.
0.8
:
0.6
r
0.4 r
0.2
7
0.8
0.9
1
v/v
1.1
1.2
305
ll.E. Absorption bands in reflectance
similar to a transmission spectrum in shape. This case is illustrated in Figure 11.7a. The reflectance decreases from wing to center and then increases back to wing again, and the band center is at the same wavelength as in transmission. If the reflectance of a monominerallic powder is transformed to an espat curve, the espat will be directly proportional to the absorption coefficient. Thus, if the band shape is Gaussian in a, it will also be Gaussian in W. If the value of a(D) at the center of the band lies in the weak surface-scattering region, then the reflectance saturates as the wavelength enters that region. This case is illustrated in Figure 11.7b. A further increase in k or a does not cause a corresponding decrease in reflectance. Instead, the reflectance remains constant until a(D) moves out of the weak surface region back into the volume-scattering region, and the reflectance increases. In this case the bottom of the band is cut off and replaced by a flat line. The heights of the wings are increased by multiple scattering effects, relative to the wings seen in transmission. A naive observer might easily interpret the band as two unresolved, overlapping bands. If k > 0.1 anywhere, then the band center will lie in the strong surface-scattering regime. This case is illustrated in Figure 11.7c. As k increases, the reflectance first decreases through the volumescattering region, but then increases as it enters the strong surface Figure 11.7b. Reflectance spectrum of a particulate material with an absorption band of Lorentzian shape whose center lies in the weak surface-scattering region. The center of the band is located at vQ.
0.8
0.6
0.4
0.2
r
0.8
0.9
1
v/v
1.1
1.2
11. Reflectance spectroscopy
306
region. The reflectance goes through a maximum near the center of the band; however, anomalous dispersion effects cause n as well as k to be functions of frequency, so that the position of the maximum of the reflectance curve lies slightly to the higher-frequency (shorterwavelength) side of the actual band center. As k decreases again, the reflectance decreases through a second minimum and then increases into the wing. Thus, a strong band has two minima on either side of the maximum corresponding to the band center. This type of minimum will be called a transition minimum, because it occurs in the weak surfacescattering transition region between the volume-scattering and strong surface-scattering regions. Anomalous dispersion causes the higherfrequency transition minimum to be deeper than the lower-frequency one. A naive observer could easily mistake this band for two overlapping, partially resolved bands. The unusual shapes of absorption bands when the band center is dominated by surface scattering means that automated methods for band identification, such as the one proposed by Huguenin and Jones (1986), must be used with great care when strong bands are present. To illustrate these concepts, let us turn to a material of practical interest in remote sensing: water ice. The complex refractive index of H 2 O ice as a function of wavelength is shown as the solid and dashed Figure 11.7c. Reflectance spectrum of a particulate material with an absorption band of Lorentzian shape whose center lies in the strong surface-scattering region. The center of the band is at v0.
0.8
0.6
0.4
0.2
0.8
0.9
1
v/v
1.1
1.2
ll.E. Absorption bands in reflectance
307
lines in Figure 11.8a. The reflectance spectrum of an H 2 O frost, measured over the same wavelength range, is shown as the solid line in Figure 11.8b. The absorption spectrum shows a strong fundamental vibrational band at 3.08 /xm, plus a number of overtone bands whose strength decreases with decreasing wavelength in the near infrared. It also shows part of a strong electron-excitation band in the vacuum ultraviolet. In reflectance, the infrared fundamental band is expressed as a maximum at 3.15 ^m, with two associated transition minima at 2.85 and 3.45. The 2.85 minimum is lower than the 3.45 minimum. The weaker remaining bands are all in the volume-scattering regime and are expressed as minima. The refractive index of H 2 O ice was inserted into equations (11.14) and (11.15), assuming 4 = 0, and the diffusive-reflectance spectrum ro(A) was calculated. Because the particle size of the sample in Figure 11.8b was not known, (D) was taken to be an unknown parameter. The value that gave the best fit was
-1
'
•
1
1
i
'
-
k/
-2 —-
/ \
-4 " -5 ~ -6
1
1
+
_
%
,'
rI
\\ \
N
~t -
-7
jr
1
\
/AV >^
-
-t
\ \
1f
-
1
c
/
_
-3
|
Jn
1 1 \ -
_
•• + i
i
i
i
i
1
1
,
I
I
1
1
1
j
i
i
i
|
1
I
,
,
.
308
11. Reflectance spectroscopy
was assumed to be equal to ro(A), and equation (11.4) was used to calculate y(A) and w(\) from the reflectance. The real part of the index of refraction was assumed to be constant at n = 1.3, independent of wavelength. (D) was assumed to be 125 jinn, and equation (11.19) was used to calculate a. From the dispersion relation, k was then found. The result is shown as the points in Figure 11.8a. The agreement is again seen to be excellent, except at two wavelengths: the strong fundamental band, and in the visible. The reason for the less satisfactory agreement at 3 /tm is because n was assumed to be constant, and anomalous dispersion was neglected. The recovery of n and k in this region could be improved by using the Kramers-Kronig relations (Section 4.D). The reason for poor agreement in the visible is the large measurement error when the reflectance is close to 1, as discussed in Section ll.D.2.b. 11.E.2. Dependence of band depth on geometry
Thus far, the shapes of the absorption bands have been discussed in terms of r 0 , which in its simplest interpretation is independent of Figure 11.8b. Spectral reflectance of water frost. Solid line; measured reflectance; crosses, diffusive reflectance calculated from the measured refractive index shown in Figure 11.10a, assuming (D) = 125 /im. The arrows show the two transition minima on either side of the band center. Data from Smythe (1975) and Hapke et al. (1981).
0.2 -
Mum)
ll.E. Absorption bands in reflectance
309
angle. However, for physically real cases, the band depths and shapes depend on illumination and viewing geometry, as has been emphasized by Veverka and his co-workers (Veverka et al, 1978a-c). If the bidirectional reflectance of a material is measured with I and e close to normal, then multiple scattering will be important in the wings of the band, where w is high, but will be less important near the band center, where w is low and single scattering is the primary contributor to the brightness. However, if the reflectance is measured at large values of I and e, only single scattering contributes throughout the whole band. Similar effects can cause differences in band shapes between spectra of the same material measured bidirectionally and using an integrating sphere to measure the directional-hemispherical reflectance (Gradie and Veverka, 1982). The dependence of the relative band shape on angle and type of reflectance is illustrated in Figure 11.9. 11.E.3. Dependence of band contrast on particle size
A question of interest is the relation between the depth of an absorption band seen in reflectance and the particle size of the scattering medium. The question may be addressed by using equations Figure 11.9. Reflectance spectra of a participate medium of isotropic scatterers with a Gaussian absorption band, illustrating the effects of illuminating and viewing geometry on the band depth. 1.2
0.8
'. ;
: -
\
bidirectional (i = 0, e = 0)
reflectance \
v/
'^X \
/
hemispherical reflectance (i = 0)
tr ©
I
0.4
DC 0.2
:
^S/
- bidirectional • (i
-
reflectance-^^^
^60 o,e^60^) 0.5
1.5
TJX
310
11. Reflectance spectroscopy
(11.14) and (11.15) for w to calculate the diffusive reflectance r0. This has been done in Figure 11.10, where for purposes of illustration we have taken n = 1.50 and 4 = 0. It was assumed that the absorption coefficient at the center of the band was aB = 1.20ac, where ac is the value in the continuum. Figure 11.10 shows the relative band contrast plotted against in reflectance A r / r = [r o (a c ) - ro(aB)]/ro(ac) ac(D). The continuum reflectance ro(ac) is also shown. When ac(D) is very small and the particles are optically thin, the band contrast is small also. As the particle size, described by (D), increases, A r / r increases to a maximum value roughly equal to the relative band contrast in absorbance, 20% at ac(D)^l. The reflectance decreases montonically. As the particle size continues to increase, the band contrast now decreases and becomes small as the particles become optically thick, and the reflectance saturates in the weak surface-scattering region. Note that the band contrast is not a monotonic function of the reflectance or of the particle size. ll.F. Effects of intimate mixtures on reflectance spectra
Using the methods of Section 10.F, the reflectance of an intimate mixture can be calculated from the reflectances of the pure end members that make up the mixture. First, vv. and pig) are found by Figure 11.10. Relation between band contrast and particle size. Solid line, relative band contrast; dashed line, continuum reflectance. The relative contrast in absorbance is 20%. 0.2
0.15
0.75
Ar/r 0.5
0.25
T
0.1
0.05
11.F. Reflectance spectra of intimate mixtures
311
inverting the appropriate equation for the reflectance r^ of the /th end member. Then w and p(g) for the mixture are calculated using the mixing formulas and are inserted into the reflectance equation to calculate the reflectance of the mixture. Conversely, if the identities and spectra of the individual members of a mixture are known, the weights N-a^QEJ/ a jr.QE^/p^D- in the mixture can be found by trial and error by fitting calculated spectra to the measured spectrum of the mixture. Deconvolutions of mixtures to find the fractions of the end members have been done in laboratory investigations by Smith, Johnson, and Adams (1985) and Mustard and Pieters (1987, 1989). Johnson, Smith, and Adams (1985) applied these concepts to the analysis of lunar reflectance spectra. Figure 11.11, from the paper by Mustard and Pieters (1989), compares the fractions of the end members in binary and ternary mixtures calculated by deconvolution of the reflectance spectra with the actual values. They deconvolved the data in two ways, one assuming that all particles scatter isotropically, and the other allowing for nonisotropic scattering. As might be expected, the nonisotropic fit gave smaller residuals. However, making the simplifying assumption that the scatterers are isotropic still allows the abundances to be estimated to better than 7%, except for opaque minerals, for which the errors are somewhat larger. The following example illustrates the mixing equations. Suppose we have two powders consisting of isotropic scatterers larger than the wavelength. Suppose that the hemispherical reflectances of the two materials, measured at I = 60°, are rhl = 0.05 and rh2 = 0.90. What will the reflectances of various mixtures of the two powders be? When I = 60° the hemispherical reflectance is equal to the diffusive reflectance. Thus, from equation (11.4), the single-scattering albedos corresponding to these reflectances are calculated to be wx = 0.181 and w2 = 0.997. For a binary mixture of particles, equation (10.61) is wx + &w2 w= 1+g where =^i
Jt2
^
P2 Pi Pl
Dx '
is the bulk density of material of type / , p. is its solid density, and
312 11. Reflectance spectroscopy D is its size. For simplicity, assume that the solid densities pt and p 2 are approximately equal. Then the weighting factor is ^ = (jf1/j?2)(D2/D1), which depends on both the ratio of particle sizes and the relative amounts of the two materials. Figure 11.12 illustrates the dependence of the reflectance of the mixture on the mass mixing ratio JK2/{JKX + **T) ^ or three different size ratios, D2/D1 = 0.01, 1.0, and 100. This figure makes the important point that when there is large disparity in the sizes of the Figure 11.11. Determination of mass fractions in intimate mixtures by deconvolution of reflectance spectra. Top, ternary mixtures of olivine, enstatite, and anorthite; bottom, binary mixtures of olivine and magnetite. The filled circles are the actual mass fractions; the open squares are the results of deconvolution assuming isotropic scattering; the open circles are the results of deconvolution assuming nonisotropic scattering. From Mustard and Pieters (1989); copyright by the American Geophysical Union.
EN
AN 0.8
o* D
0.6
I
"5
O
0.4
D
0.2 -
o
-
D
#
Q«
0 1 0.2
1
1 0.4
1
1 0.6
1
1 0.8
Mass Fraction of Olivine
1
1
1.0
11.F. Reflectance spectra of intimate mixtures
313
components of an intimate mixture, the fine particles can have an effect all out of proportion to their mass fraction. For example, Clark and his co-workers (Clark and Roush, 1984; Clark and Lucey, 1984) found that the addition of a small amount of finely divided carbon black to coarse particles of ice drastically reduced the reflectance and the contrast in the absorption bands of the ice. Note that if the bright material in the example in Figure 11.12 possessed any absorption bands, they would be almost totally masked by only a small amount of the dark material. The same effect also accounts for the low albedo of the moon. Lunar regolith consists of pulverized rocks and glasses of the same composition as the rocks. If a lunar rock and a glass made by melting the rock in vacuum are finely ground, the resulting powders are found to have a much higher albedo than the soil (Wells and Hapke, 1977). However, in the regolith many of the rock and glass fragments are welded together into particles called agglutinates, which are quite dark because they also contain submicroscopic particles of metallic iron. About 0.5% of the soil consists of this submicroscopic metallic iron, an amount that is sufficient to lower the reflectance of the mixture to the observed value (Hapke, Cassidy, and Wells, 1975). Figure 11.12. Diffusive reflectances of intimate mixtures of bright and dark particles as a function of the mass ratio for three different particle size ratios. Note that when the low-albedo particles are much smaller than the high-albedo particles, the reflectance of the mixture is almost independent of the amount or the reflectance of the bright material.
0.8
0.6
0.4
0.2
0.2
0.4
0.6
0.8
314
11. Reflectance spectroscopy
It is frequently stated in the literature that in a mixture of large and small particles, the small particles "coat" the large particles and prevent light from reaching them, so that the large particles cannot influence the reflectance. This is a physically incorrect explanation, because the effect occurs even when the particles are so far apart they never touch. Small particles have a large influence because of the combined effects of the nonlinear dependence of reflectance on single-scattering albedo plus the weighting of the properties of the components by cross-sectional area rather than volume. ll.G. Absorption bands in layered media ll.G.l.
The effect of layers on band contrast
Always in the laboratory and often in the field we deal with layered media. Thus, an important question of practical interest is how layers affect our ability to detect and measure diagnostic absorption bands that may be present. A band may be displayed by the upper layer, in which case we wish to know how thick the layer must be for the band to be visible or well developed. Conversely, the band may be in the lower layer, in which case we wish to know how thin the upper layer must be in order not to hide the band. As might be expected, the answers to these questions depend on the scattering properties of both of the layers. To be rigorous, these questions should be addressed using the two-layer bidirectional equations developed in Section 9.D. However, for a semiquantitative discussion, the two-layer diffusive model, equation (9.14), may be used. Define the band contrast of a layered medium as
where rw is the reflectance in the wing of the band, and rc is the reflectance at the center. This contrast will be a function of the optical thickness T 0 of the upper layer. Define the relative band contrast A C ( T 0 ) as
That is, AC is the ratio of the band contrast observed in a layered medium to the intrinsic contrast the band would have if the medium
ll.G. Absorption bands in layered media
315
exhibiting it were infinitely thick and not covered by any other material. To illustrate the effects of layering on band contrast, we will consider four examples, in which it is assumed that one of the layers has a band with an intrinsic contrast of 20%. Example 1: band in top layer, top layer dark, bottom layer bright Suppose the bottom layer has a reflectance r^ = 0.90, and that in the wings of the band the top layer has r<% = 0.090, with the corresponding y = 0.83. In the center of the band, r<% = 0.072, with the corresponding y = 0.87. Using equation (9.14), the reflectance r0 and relative band contrast AC may be calculated as functions of the optical thickness T 0 of the upper layer. The curve of reflectance is plotted in Figure 11.13 (top), and that of relative contrast in Figure 11.13 (middle). As T 0 increases, r0 decreases, while AC increases rapidly. Both reach approximately their thick-layer values by r 0 - 2. Hence, a relatively thin layer is all that is required to develop both the reflectance and the band fully. Example 2: band in top layer, top layer bright, bottom layer dark. Suppose the bottom layer has r%, = 0.090; in the wings of the band the top layer has r<% = 0.90, with corresponding y = 0.053; in the band center, r%, = 0.72 and y = 0.16. The curves of reflectance and relative contrast for this case are plotted in Figures 11.13 (top) and 11.13 (middle), respectively. In this case both the reflectance and the relative contrast change much more slowly as T 0 increases and have not reached their full values even for an optical thickness of r 0 > 11. The multiply scattered light penetrates deeply into the medium, so that the lower layer influences the reflectance even through an optically thick layer. Example 3: band in bottom layer, bottom layer dark, top layer bright In this case the band is in the bottom layer. Suppose that in the wings of the band the reflectance of the bottom layer is r# = 0.90, and in the band center it is r%, = 0.72. Let the upper layer have ry = 0.90, with corresponding y = 0.83. The reflectance and relative contrast are shown in Figures 11.13 (top) and 11.13 (bottom), respectively. Although the reflectance increases slowly with increasing r 0 , the relative contrast decreases rapidly, and an optical thickness of only r 0 ~ 1 is sufficient to hide the band almost completely.
316
11. Reflectance spectroscopy
Example 4: band in bottom layer, bottom layer bright, top layer dark. In
the wings of the band, let the reflectance of the bottom layer be A> = 0.09, and at the band center, r^ = 0.072. Suppose the upper layer has r^ = 0.90 and y = 0.053. The corresponding curves are plotted in Figures 11.13 (top) and 11.13 (bottom). Both the reflectance Figure 11.13. Reflectance and relative absorption-band contrast in twolayer media showing the effects of the reflectances of the top and bottom layers and of the layer in which the band is located. (Top) Diffusive reflectance, ru = 0.9, rL = 0.09. (Middle) Band contrast, band in upper layer, (Ar/r)^ = 20%. (Bottom) Band contrast, band in lower layer, (Ar/r) L = 20%. 0.8
bright over dark
_I
r L 1
0.6
0.4
-
w -/
-
0.2
dark over bright
I
.
I
.
i
.
i
.
i
.
i
.
i
.
i
.
10
I dark over bright 0.8
bright over dark
0.6
0.4
0.2 Absorption Band in Top Layer
10
11. G. Absorption bands in layered media
317
and relative contrast decrease rapidly with increasing r 0 , and the band is practically invisible when T0 > 2. 11.G.2. The radialith, or How thick is "thick enough"?
A question of interest in many remote-sensing applications is, How thick is the layer that controls the amount of light reflected from a planetary regolith? Nash (1983) has termed this layer the radialith. This question may be answered using the expression for the two-layer diffusive reflectance, equation (9.14). In that expression the terms that contain the scattering properties of the lower layer are proportional to exp( - 4yT 0 ), where y = Jl-w is the albedo factor of the upper layer, and T 0 is its optical thickness. Thus, the effects of the lower layer will be reduced to a small value if exp(-4yT 0 )<^l. This suggests a convenient criterion for the thickness of the radialith as that necessary to make 4yr 0 = 6. If the density of particles is uniform in the medium, T0 = NaQEzR, where N is the number of particles per unit volume, or is their mean cross-sectional area, QE is their extinction efficiency, and zR is the thickness of the radialith. For large particles, QE = 1, and we may write, approximately, (11.24) Figure 11.13. (cont.)
0.8
dark over bright
0.6
0.4
bright over dark
0.2 Absorption Band in Bottom Layer
~
10
318
11. Reflectance spectroscopy
where
.
(11.25)
If w is small, then y is not too different from 1, and if the particles are close together, so that
A
' ~
(11.26)
If the particles are very weakly absorbing, and if, in addition, the porosity of the layer is high, the radialith can be very thick indeed. The same criterion may be used to estimate the thickness required for laboratory samples in order that the substrate not influence the reflectance. If the particles are absorbing over the wavelength range of interest, then only a few monolayers are required. If the absorbance is very small, then equation (9.16) may be used to estimate the necessary thickness, Suppose our criterion is that r0 change by less than 1% no matter what the reflectance of the lower substrate r^. This requires T 0 > 99, which from (11.24) means that the layer must be more than 66/<£ particles thick. If the filling factor is of the order of \, the sample must contain more than 130 monolayers. For example, if the grain size is of the order of 80 ju,m, the sample must be at least 1 cm thick. 11JL Kubelka-Munk theory: What's wrong with it? 11.H.1. Derivation of the Kubelka-Munk equations
A model that is widely used to interpret reflectance data is a version of the diffusive reflectance known as Kubelka-Munk theory (Kubelka and Munk, 1931). Unfortunately, this model contains several inherent limitations that greatly restrict its ability to correctly describe the reflectance of particulate media. In this section the Kubelka-Munk equations will be derived and analyzed critically in order to clarify the nature of the limitations. All parameters in Kubelka-Munk theory will
ILK Kubelka-Munk theory: What's wrong with it? be denoted by primes in order to distinguish them from similar parameters that appear in the diffusive-reflectance equations derived previously. The Kubelka-Munk model begins with the equation of radiative transfer (Chapter 7), which states that the divergence of the radiance as it propagates through a volume element in a particulate medium is equal to the sum of three terms: the partial extinction of the wave, the radiance traveling in other directions that is scattered into the direction of propagation of interest, and the addition from source terms. The Kubelka-Munk model is a two-stream type of solution to the radiative-transfer equation. As usual, denote the radiances traveling into the upward and downward directions by Ix and / 2 , respectively. Then their divergences are \{dlx/dz) and —\{dI 2/dz), respectively. The terms \ and — \ arise from averaging the cosine of the direction of propagation over the upward and downward hemispheres (see Section 8.F). The extinction of the radiance is by absorption and scattering. Denote the Kubelka-Munk volume absorption coefficient by K', and the volume scattering coefficient by 5'. Then the extinction coefficient is (K' + 5"). In Kubelka-Munk theory, K' is assumed to be equal to ( a ) , the true absorption coefficient a inside the particles of the medium averaged over a unit volume, and the scattering is assumed to be due to undefined scattering centers that are entirely independent of the absorption. The radiance is assumed to be scattered only if its direction of propagation has been changed from the hemisphere into which it was moving to the oppositely going hemisphere. Light that is scattered into the same hemisphere is interpreted as unscattered. As in the diffusive reflectance, the volume source term is zero. The only source is a uniform diffuse radiance emerging into the downward hemisphere from the invisible "membrane" covering the upper surface of the medium, and it appears in the boundary condition at the surface. With these assumptions, the equations governing the radiance consist of two coupled differential equations,
^
= -(K' + S')I1 + S'I2,
(11.27a)
-\^£
= -{K' + S')I2 + S>Iv
(11.27b)
319
320
11, Reflectance spectroscopy
Let
dr' = -(K'+2Sf)dz,
(11.28a)
and w' = K,+2S' •
(H.28b)
Then (11.27) can be put into the form
-lW
=
-^ + Y^
+I
^
(1L29a)
W = - 7 2 + T ^ + / i)(1L29b) Comparing (11.29) with equations (8.17) we see that they have exactly the same form. Because the boundary conditions are also the same, we may write down the expression for the Kubelka-Munk reflectance by comparison with (8.25),
(11.30) where y' = Vl — w'. Solving TQ for K'/Sr gives the so-called Kubelka-Munk remission function, f(r'0\
By assumption, K' is identified with the volume-averaged absorption coefficient (a), and S' is independent of (a); thus, / ( ^ Q ) should be proportional to , with the constant of proportionality equal to 1/5'. In particular, the true absorption coefficient a of a monominerallic medium should be proportional to /?(r'o). Experimentally, it is found that if f(r'o) is calculated from a measured directionalhemispherical reflectance, while a is independently measured by transmission, the two quantities are indeed proportional to each other for small values of f(rr0). However, with increasing absorption, the slope of the remission function decreases, and the curve saturates. There has been a great deal of discussion in the literature as to the reason for this failure at larger absorptions. The usual explanation is that the scattering coefficient 5' is somehow "wavelength-dependent."
ILK
Kubelka-Munk theory: What's wrong with it?
321
Because the proportionality between a and f(r'o) is valid only for small values of the remission function, a technique has been developed called the dilution method, which attempts to measure larger values of a by reflectance: A few of the strongly absorbing particles are mixed with a sufficient quantity of weakly absorbing particles to bring the mixture into the linear region. However, although this somewhat extends the linear range, in practice the amount of improvement is found to be minor.
11.H.2. A critique of Kubelka-Munk theory
Let us analyze Kubelka-Munk theory to see where the difficulties are. We will discuss the problems in order of increasing severity. (1) Radiance that is scattered in the same hemisphere into which it was originally traveling is assumed to be unscattered. This is incorrect in principle. However, in practice it means that 5' should really be replaced by S'/2. But because 5' is nebulously defined, this does not lead to serious errors of interpretation of experimental data. (2) The Kubelka-Munk reflectance is a form of the diffusive reflectance, which has no physical meaning because of the nonphysical boundary conditions. It is used to interpret measured reflectances that are usually either directional-hemispherical or bidirectional. However, we have seen in previous portions of this book that if properly used, the diffusive reflectance is in fact capable of giving quantitatively correct results. Hence, the use of this formalism to reduce data may or may not result in serious errors, depending on the conditions under which the reflectance was measured. (3) S' is nebulously defined and cannot be predicted from properties of the material, such as refractive index and particle size. However, several authors have noted that S' is of the order of the reciprocal of the particle size. (4) The parameters K' and S' are completely misinterpreted. This is the most serious error in Kubelka-Munk theory and is the source of the linearity failure in the remission function and the dilution method. The problem is that the absorption and scattering are not distributed evenly throughout the medium, as assumed by the theory, but are localized into particles. If the Kubelka-Munk differential equations are compared with the diffusive-reflectance
322
11. Reflectance spectroscopy equations for a particulate medium, then K' is seen to be equivalent to K, and 5' to 5/2. Thus, Kr is not equal to the average internal absorption coefficient of the particles, but is the volume absorption coefficient K = %.Nj(TjQA., which is an entirely different quantity. Similarly, 5' = 5 / 2 = 2<:N:(T-QS;/2. Furthermore, because for large particles Qs; + QA ; = QE • = 1, 5' is not independent of Kf, but has a behavior complementary to that of K': 5' = X^N^aXl — QAj)/2 = (5^A^ov — K')/2. For large particles, the reason that 5' is not constant has nothing to do with wavelength-dependent scattering, but occurs because 5' and K' both depend on the absorption. 11.H.3. Relation of the remission Junction to the espat Junction and the limitations of the dilution method
From equation (11.4b), the single-scattering albedo is related to the diffusive reflectance by w = 4 r o / ( l + r 0 ) 2 . Hence, for the diffusive reflectance the espat function is given by
Comparing (11.32) with (11.31) shows that the remission function is equal to twice the espat function, the factor of 2 arising from the different definitions of 5 and 5'. We can now understand why the remission function behaves as it does. It is equal to twice the espat function and hence is subject to the same limitations as W. It is proportional to the internal absorption coefficient of a monominerallic material if the reflectance is in the volume-scattering region, but it saturates as the reflectance enters the weak surface-scattering region. Even though Kubelka-Munk theory misinterprets the physical nature of the remission function, f(r'o) turns out to be proportional to the true particle absorption coefficient in the volume-scattering region because of the fortunate mathematical coincidence in the behaviors of QA and Qs. This is why the theory is useful. We can also understand why the dilution method does not appreciably improve the linearity. It was shown in Section ll.D that the weighting factors of a> in the espat of a mixture are independent of a- only if a^De^<:l for all components. Suppose a small amount of material with a high absorption, denoted by subscript 2, is mixed with
ILK Kubelka-Munk theory: What's wrong with it?
a large amount of very weakly absorbing material, denoted by subscript 1. Then the espat function of the mixture is, from (11.21),
where & = N2
323
324
11. Reflectance spectroscopy
processing. A more accurate procedure would be to calculate the absorption coefficient or the espat function using one of the equations appropriate to the geometry of the instrument given in this chapter. The dilution method offers no substantial advantage in extending the linear region of the remission function. Instead, it is recommended that the particles be ground to a small enough size that the espat function will be in the volume-scattering region; that is, aD < 3 for the largest anticipated value of a. If k is as large as 0.1, this means that D < 2.5A. Larger values of k will place the reflectance in the strong surface-scattering region and allow the retrieval of k by Kramers-Kronig analysis (Section 4.D).
12 Photometric effects of large-scale roughness
12.A. Introduction
The expressions for reflectance developed in previous chapters of this book implicitly assume that the apparent surface of the particulate medium is smooth on scales large compared with the particle size. Although that assumption may be valid for surfaces in the laboratory, it is certainly not the case for planetary regoliths. In this chapter the expressions that were derived in Chapters 8-10 to describe the light scattered from a planet with a smooth surface will be modified so as to be applicable to a planet with large-scale roughness. In calculations of this type we are immediately faced with the problem of choosing an appropriate geometric model to describe roughness. Some authors have chosen specific shapes, such as hemispherical cups (Van Diggelen, 1959; Hameen-Anttila, 1967), that approximate impact craters on the surface of a planet. However, such models may not be applicable to other geometries. To make the expressions to be derived as general as possible, it will be assumed that the surfaces are randomly rough. There is a large body of literature that treats shadowing on such surfaces - see, for example, Muhleman (1964), Wagner (1967), Saunders (1967), Hagfors (1968), Lumme and Bowell (1981), and Simpson and Tyler (1982), as well as the references cited in those papers - although many of those papers deal only with specular reflection, such as is involved in analyses of sea glitter or backscattered lunar radar signals. In order to treat diffuse bidirectional-reflectance functions, the approach of Hapke 325
326
12, Photometric effects of large-scale roughness
(1984) will be followed. As in the other parts of this book, the emphasis will be on the development of useful approximate analytic expressions, rather than perfect mathematical rigor. The derivation is based on the following assumptions: (1) Geometric optics is valid. If the medium is composed of particles smaller than the wavelength, the objects that control the scattering are large clumps rather than individual particles. (2) The macroscopically rough surface is considered to be made up of small, locally smooth facets that are large compared with the mean particle size and are tilted at a variety of angles. The normals to the facets are described by a two-dimensional slope distribution function aidtOdftdC, where # is the zenith angle between a facet normal and the vertical, and £ is the azimuth angle of the facet normal. It will be assumed that the distribution function of the facet orientations is independent of azimuth angle, so that &(#,£) can be written simply as aid). Then, in general (Hagfors, 1968), if d(ft) is the one-dimensional function that describes the distribution of slopes on any vertical cut through the surface made at an arbitrary azimuth angle, the corresponding two-dimensional, azimuth-independent distribution function can be written in the form (12.1) It will be assumed that &'(#) can be described by a Gaussian distribution of the form a (#) dd = j/exp[ - 3§tan2#] rf(tan #),
(12.2)
where J / and 38 are constants to be determined. Then, a,(#) = j / e x p [ - ^ t a n 2 # ] s e c 2 # s i n # .
(12.3)
The slope distribution function is normalized such that )
l
(12.4)
o
and is characterized by a mean slope angle 0 defined by tan0 = - r / a(#)tan#d#.
(12.5)
Inserting (12.3) into (12.4) and (12.5) shows that ,
(12.6a)
12.A. Introduction
327
and ^ = l/7Ttan 2 0.
(12.6b)
An important question is the physical meaning of 0. This parameter is the mean slope angle averaged according to (12.5) over all distances on the surface between upper and lower limits that are determined by the angular resolution of the detector and the physics of the radiative-transfer equation. The upper limit is the footprint of the detector on the surface of the planet, which in planetary remote sensing is typically hundreds of meters to kilometers. It might be thought that the lower limit is given by the sizes of the particles making up the surface. However, it must be remembered that the radiative-transfer equation for a particulate medium, on which the solutions for the reflectance are based, implicitly averages the radiance over distances that are large compared with the distances between the particles. Thus, the lower limit is several times the mean particle separation. A typical size for this lower limit is of the order of a millimeter. Moreover, the maximum slopes that can occur on natural surfaces are determined by the strengths of materials and the cohesiveness of the soil. The effects of these properties are strongly size-dependent, such that the small-scale slopes tend to be the highest. Thus, the slope distribution functions tend to be dominated by millimeter-scale roughness. The assumption that the slope distribution function is independent of azimuth will certainly be true on the average for surfaces made up of craters and hills, and it also appears to be reasonable for the surface of the ocean (Cox and Munk, 1954). Its validity may be questioned for morphologies with preferred orientations, like folded mountain ranges and fields of parallel sand dunes. However, on the small scales that dominate the distribution function, the slopes are likely to be caused by such erosive agents as microscopic impacts, eolian gusting, and fluvial action, which are roughly isotropic in azimuth. Hence, the assumption appears to be reasonable. (3) The mean slope angle 0 is assumed to be small. Vertical scarps and overhangs are assumed to constitute a negligible part of the surface. Although the general equations for the roughness effects will be derived for an arbitrary 0, the analytic expressions are greatly simplified if terms of order 03 and higher can be ignored.
328
12. Photometric effects of large-scale roughness
(4) It is assumed that light multiply scattered from one surface facet to another is negligible. However, radiance that is multiply scattered from one particle to another within each surface facet is included in the derivation. The limitation that this assumption imposes can be estimated by the following calculation: Consider a depression in the shape of a sector of a sphere of radius x and with maximum slope # M . The inside is covered with a Lambert surface with albedo AL. The depression is illuminated vertically by irradiance / (see Figure 12.1). Then the radiance scattered once from a small area A A at the bottom of the cup is
What is the radiance I2 due to light scattered onto A A from the rest of the inside of the cup? Consider an increment of area dA = x2 sm&d&dif/ located at zenith angle # and azimuth t// on the inside of the cup. Then the doubly scattered radiance from A A due to light scattered from dA is dl2 = —dA cos 2
77
j cos & L
Figure 12.1.
J
IT
12.B. Derivation
329
where &' and y are defined in Figure 12.1. Now, #' = (ir - #)/2, and y = 2xsin(#/2). Hence, the radiance from kA due to light scattered from the entire inside of the cup is •'^ = 0*^^ = 0
^
4TT
and the ratio of doubly to singly scattered radiance is
For example, if AL = 0.5 and &M = 45°, I2/Ix = f>%. The assumption that interfacet scattering can be neglected is seen to be consistent with assumption (3). Light multiply scattered from one facet to another usually can be ignored if either the albedo or mean slope is small. The major exception to this statement occurs for high-albedo surfaces at large phase angles. Then most of the visible facets may be in the shadow of the direct irradiance from the source, but will still be illuminated by light scattered from adjacent surfaces not in shadow. In order to account for interfacet scattering, Lumme and Bowell (1981) suggest applying a roughness correction only to the singlescattering term in the equation for the reflectance of the individual facets, but not to the multiple-scattering term. However, this assumption has no physical justification and seems to be a case of attempting to make a right by compounding two wrongs. The approach preferred in this book will be to recognize that most bodies in the solar system have fairly low albedos, so that the neglect of interfacet scattering is reasonable; however, the model must be applied with care to those outer-planet satellites with high albedos. 12.B. Derivation 12.B.1. Derivation of the general equations
The general scheme of the derivation will be as follows. We will seek a formalism by which the bidirectional reflectance of a medium having a smooth surface can be corrected to one describing the same medium, but with a surface roughness characterized by a mean slope angle 0. General equations that are mathematically rigorous will be derived first, and the parameters necessary for their evaluation will be defined. Because the effects of roughness are maximum at grazing illumination
330
12. Photometric effects of large-scale roughness
and viewing, these expressions will be evaluated to obtain analytic functions that are exact for these conditions. Next, the equations will be evaluated for vertical viewing and illumination. The two solutions will be connected by analytic interpolation to give approximate expressions for intermediate angles. Consider a detector that views a surface having unresolved roughness from a large distance <^ and that accepts light from within a small solid angle Ao> about a direction with zenith angle e. The signal I(i,e,g) from this detector is interpreted as if it came from a smooth, horizontal area A = c^2Aa>sec e on the mean surface with bidirectional reflectance rR(l,e,g); that is, ,
(12.7)
where / is the incident irradiance. The model assumes that the light actually comes from a large number of unresolved facets that are tilted in a variety of directions and are both directly illuminated by light from the source and visible to the detector. Let the bidirectional reflectance of each individual facet be r(l, e,g), and let each facet have area Af <^c A. The geometry is shown schematically in Figure 12.2. Then the true expression for Figure 12.2. Schematic diagram of the intersection of the surface and a vertical plane containing the detector. Shown are the actual surface, consisting of a multitude of unresolved facets Af, the nominal surface A, and the effective tilted surface A t. A cut by a vertical plane containing the source would be similar.
12.B. Derivation
331
the light reaching the detector is [
r(it,6t,g)ca&etdAt9
AU,v)
(12.8)
where the subscript t (standing for "tilted") on I, e, and A denotes values appropriate to an incremental surface of area (12.9a) whose normal points in direction (#,£), and the symbol (i,v) indicates
that the integration is to be taken only over those surface facets within
A that are both illuminated and visible. Using the law of cosines it may be readily shown that the angles it, en # , £, and */f, where $ is the azimuth angle between the source and detector planes, are related by cos it = cos I cos ft + sin I sin # cos £, cos tt = cos ecos# +sin esin#cos(£ — ijj).
(12.9b) (12.9c)
Note that g is the same for all facets within the area viewed by the detector. The objective of this chapter is to find the relation between r(l, e,g) and rR(i, e,g) as a function of 0. There are three important effects of macroscopic roughness that will modify the reflectance: (1) Scattering of light from one facet to another will increase the reflectance. This effect will be small if either the albedo or the mean slope is small, as argued in the preceding section, and will be ignored. (2) Unresolved shadows cast on one part of the surface by another will decrease the reflectance. (3) As the surface is viewed and illuminated at increasing zenith angles, the facets that are tilted away from the observer or source will tend to be hidden or in shadow, so that the surfaces that are visible and illuminated will tend to be those that are tilted preferentially toward the detector or source. To account for the latter two effects, we will try to write the rough-surface bidirectional reflectance rR(i, e,g) as the product of a shadowing function S(£, e , g ) and the bidirectional reflectance r( ie9 ee, g) of a smooth surface of effective area Ae tilted so as to have effective angle of incidence le and angle of emergence e e , and with the same phase angle g. That is, we will seek expressions for ie(i, e, g), £*(£> t9 g\ a n d SG, e,g) that will make the following equation true: rR(i, e , g )= r(Le, ee,g)S(i, e , g ) .
(12.10)
332
12. Photometric effects of large-scale roughness
From Figure 12.2, A and Ae are related by Aecosee = Acose. Let (12.12a) (12.12b) (12.12c) cosi r
(12.12d)
Denote the reflectance per unit area of surface by Y; that is, let YR(fi0,/x,g) = rR(i,e,g)cos6, >V<e>8) = r(ie, ee,g)cos te, t,ixt,g)
= r(Lt,et,g)coser
(12.13a) (12.13b) (12.13c)
Combining (12.7)-(12.9),
= jf J
(12.14)
AU,v)
Assume that Y is mathematically well behaved so that it can be expanded in a Taylor series about fi0 and fi. Doing this on both sides of the third equal sign in (12.14), and using (12.11), gives Afju
dY dY
AU,v
f.
dY
A(i,vY •jy
12.B. Derivation
333
or
t
(
t
)
t
+ - ~ .
(12.15)
Because /x0 and JJL are independent variables and Y can be an arbitrary function of these variables, equation (12.15) will be satisfied if the coefficients of Y and its partial derivatives are separately equal on both sides of the equality. This gives
JA{i,v)a/Lt
(12.18)
In general, there will be two types of shadows. Some of the facets will not contribute to the scattered radiance because their normals will be tilted by more than 90° to the direction from the source or detector; such facets will be said to be in a tilt shadow. Some facets will not contribute because other parts of the surface will obstruct either the view of the detector or the light from the source; such facets will be said to be in a projected shadow. We shall follow Sanders (1967) and assume that any facet that is not in a tilt shadow has a statistical probability of being in a projected shadow that is independent of the slope or azimuth angle of its tilt. Let P be the probability that a facet is not in a projected shadow. Then in equation (12.15), Y(fjLOn/jLng) can be multiplied by P, if at the same time the boundaries of the integration are replaced by the tilt-shadow boundaries. This has the effect of multiplying both the numerators and denominators in (12.17) and (12.18) by P, which thus cancels out in these equations. Therefore, inserting (12.9) into (12.17) and (12.18) and writing the latter out explicitly, these equations
334
12, Photometric effects of large-scale roughness
become cos i/(^(tiit)COS & a,(#) dftdt, + sin £/^(tilt)sin #cos £ &
(12.19) cos eX4(tilt)cos # aid) dddt; + sin e£4(tilt)sin # cos(£ - if/) ai&) dftd£
(12.20) where ^l(tilt) denotes the boundaries of the tilt shadows in A. Let AT be the total area of all the facets within the nominal area A9 whether visible and illuminated or not:
AT = f2ir r/2dAt ^ =0 ^=0
= 2TT r/2Afa(&)
d& = ITTA,,
^0
because of azimuthal symmetry and the normalization condition on a(#). Now, A is just the projection of all the facets onto the horizontal plane: 2v
/
fir /2
/2
/ 6=0^ = 0
Hence, where
(12.21)
Thus, expression (12.16) for S can be written
(12.22) where Fo,H,#)-3-f
dAt
(12.23)
is the probability that a facet is both illuminated and visible. Let F(( i) and Fe( e) be the fraction of the facets that are illuminated and the fraction that are visible, respectively. Because of azimuthal symmetry both Ft(i) and Fe(e) are independent of if/, and furthermore, Ft{l) has the same functional dependence on i as Fe(e) has on e.
12.B. Derivation
335
12.B.2. The case when i < e The solutions for S, /jbOe, and fie have different forms depending on whether I is larger or smaller than e. We will consider the case when I < e first. Then an illumination shadow cast by a given object is always smaller than its visibility shadow, and the illumination shadow may be regarded as partially hidden in the visibility shadow. When I/J = 0 the illumination shadow is completely hidden, so that F(iv)(jjLo,/jL,0) = Fe(e). B u t w h e n ^ = 0 a n d l < e , n o shadows a r e visible: The detector's field of view is completely filled by surfaces that are both visible and totally illuminated. Thus,
where H.(0)-v..(yr-0).
(12.24)
Hence, '
(
}
Because of the symmetry between Fe(e) and Ft{l),
where Moe(O)- ^ ( V ^ = 0 ) .
(12.27)
When y/±Q the visibility shadow only partially hides the illumination shadow. Let f(i//)bc the fraction of the illumination shadow that is hidden in the visibility shadow. Let
Ae/AT
= 1-Fe(e)
be the fraction of the facets in the visibility shadows, and let be the fraction of the facets in the illumination shadows. These include both the tilt and projected shadows. As we have seen, when iff = 0, all of the illumination shadows are hidden in the visibility shadows. The visibility and illumination shadows are perfectly correlated, so that /(0) = 1 and
*0»
=
*e = 1~
Ae/AT.
336
12. Photometric effects of large-scale roughness
As e/f increases, a fraction 1 - f{\fj) of the illumination shadows will be exposed. When i/r = ir, /(TT) = 0, and the two types of shadows are completely uncorrelated, so that
-(l-Ae/AT){l-Ai/AT)-FeFi9 where the term AeAt/A% corrects for the amount of random overlap. This last expression states that when the two types of shadows are completely uncorrelated, the probability that a facet will be both illuminated and visible is the product of the separate probabilities. When 0 < if/ < IT, At in the last expression for F{iv) must be replaced by (1 — f)A( /(AT — fA^. This accounts for the fact that only a portion 1 - / of the illumination shadow is randomly exposed, and only an area AT - fAt is available to be occupied by the uncorrelated part of the illumination shadow. Thus, the general expression for the probability that a facet will be both illuminated and visible is
-[l-Ae/AT][l-Ai/AT]/[l-JAi/AT]
-FA/il-f + ft). Combining this result with (12.22), (12.25), and (12.26) gives
(12.28) Thus far, the derivation has been rigorous. In the remainder of this section we will derive approximate analytic expressions for fiOe9 \xe, and S that are suitable for practical calculations. First, an expression for f(il/), the fraction of the illumination shadow hidden in the visibility shadow, will be found. It will be assumed that fit//) is a function of if/ only and is independent of i and e, which is reasonable if I and e are near 90°. Recall that the shadows have two components, tilt and projected shadows. When I and e are near 90°, the contributions of the two components are roughly equal. As if/ increases from zero, the fraction of the tilt component of the illumination shadows that are exposed increases approximately linearly with \fj> and the exposure is complete when i// = 7r. Hence, the contribution of the tilt component to / will be ~ ! ( ^ / i r ) . Now, the surface can be considered as consisting of
12.B, Derivation
337
depressions and protuberances of mean width Ad and mean height (Ad/2)tan0. When i and e are near 90°, the projected shadows are cast by objects of width ~ Ad onto surfaces a distance ~ Ad away, so that this component of the illumination shadows is nearly completely exposed when i// > 1 radian. At \fj = 1, the fraction of each type of shadow exposed is approximately f x ( l / 7 r ) + | x l = §. Hence, / ( ^ ) may be described by a function that decreases linearly from a value of /(0) = 1 to /(I) — 1 — f = j , and then decreases to zero as ^ -> TT. A simple function with the required properties is
/0A) = exp(-2tanf),
(12.29)
and this will be adopted for /(^)Returning to ixQe and ixe, only the tilt shadows affect these quantities. The boundary of the tilt illumination shadow can be found by putting it = TT/2 in (12.9a). This gives cos C = - cot # cot I.
(12.30)
This equation has no solution when 0 < # < TT/2 — £, but TT/2 < £ < 3 T T / 2 when TT/2i, <&
(12.31)
which has no solution when 0 < # .
(12.32)
If the tilt shadows are represented in a polar diagram with # as the radial variable and £ as the angular variable, then at grazing incidence and viewing the tilt-shadow boundaries are the straight radial lines £ = if/ - TT/2 and £ = TT/2, and the limits on # are 0 to TT/2. Hence ) ^ _ ' , where
(12.33)
338
12. Photometric effects of large-scale roughness
Equation (12.33) shows that when tj/ = 0 the effective surface is tilted toward the source and detector by an angle cos" 1[(2/7r)<sin^>]. For intermediate values of I and t these integrals are much more difficult to evaluate, because the shadow boundaries (12.30) and (12.31) are overlapping curves in the (#,£) polar diagram. Approximate expressions may be obtained for these integrals as follows: First the separate effects of the illumination and viewing shadows will be found. Next, the effect of overlapping the two shadows will be estimated by replacing the curved boundary lines by straight lines of constant £ and circles of constant # in the (#, £) plane. This approximation will then be improved by substituting the results from the solutions for the separate shadows. Finally, the integrals will be evaluated. The effect of the illumination shadow alone on fiOe may be seen by setting t = 0 and using (12.30) as the ^4(tilt) boundary in (12.19). Then the integration over £ in the first integral in the numerator of (12.19) may be carried out exactly to give
f
-L
ir/2-i,
0 r/2 /
i
TT/2-
2sin- 1 (l~cot 2 ^cot 2 0 1/2 cosda(^)rfd,
(12.34a)
where the value of the sin" 1 lies between TT/2 and ir. Now, the factor (1 — cot 2 #cot 2 t) 1 / 2 has the following properties: It rises with infinite slope from 0 at # = TT/2— I to 1 at # = ir/2, where it has slope 0. Thus, as a first approximation, this factor may be replaced by a unit step function at # = TT/2— L. Then equation (12.34a) becomes
f
cos# 1
2TT COS # &{p) d&+ r J
u cos # a,(d) rfi?. (12.34b)
ir/2-i
Similarly, the effect of the visibility shadow alone may be ascertained by setting / = 0 and using (12.31) in the integrals in (12.19).
12.B. Derivation
339
Then the first integral in the numerator of (12.19) is
fJr/21T/Z-
e6 2lT COS
/
2sin- 1 (l-cot 2 #cot 2 e) 1/2 cos#a(#)rf#,
+ r
(12.34c)
•V/2- e
which, upon approximating (1 -cot 2 #cot 2 e) 1 / 2 by a unit step function at # = TT/2- e, becomes
f
cos & r/2- e
rw/2
2TT COS # a(#) d& + /
ir cos t^ a(^) d&,
(12.34d)
As shown in Figure 12.3, these approximations are equivalent to replacing the curved illumination shadow boundary in the (#,£) Figure 12.3. Schematic diagram showing the square-cornered approximation to the tilt shadows in (d, £) space for the case when I <e. The projected shadows are assumed to be randomly distributed over the part of (#, £) space not in tilt shadows.
C=TI/2
340
12. Photometric effects of large-scale roughness
diagram by a square-cornered shadow bounded by the radial lines £ = TT/2 and 3TT/2 and the circles & = ir/2- I and ?r/2, and replacing the visibility shadow by a square-cornered shadow bounded by the radial lines £ = 7r/2 +
A(Xi\i)
cos # a ( # ) #£= r/2~*2TT
COS &
J
0
+ f7r/ J
7T cos &
ir/2-
t
^ 7 7
i ir/2-
J
(12.34e)
i
Now, the first two terms on the right-hand side of (12.34e) are the same as those of (12.34d). The third term on the right-hand side of (12.34e) corrects for the fraction of the illumination shadow sticking out from behind the visibility shadow and is almost identical with the last term on the right side of (12.34b), except that only a fraction t^/ir of this term contributes. But the terms in (12.34b) and (12.34d) are approximations to the exact expressions in (12.34a) and (12.34c), respectively. This suggests that the approximations in (12.38) may be improved by substituting the exact expressions from (12.34a) and (12.34c). If this is done, we obtain
[ A(tili)
= f7
/2 C
~
7T/2-C
TT/2-
i
where the value of sin~ 1 (l-cot 2 ^cot 2 e) 1/2 lies between TT/2 and TT, but the value of sin~ 1 (l-cot 2 t^cot 2 0 1/2 lies between 0 and TT/2.
72.5. Derivation
341
By an identical argument the following expression for the integral in the denominator of (12.19) is obtained: •O4(tilt)
^0
+ r/2 J
TT/2-
2sin- 1 (l-cot 2 #cot 2 e) 1/2 a(#)# e
if/ r t* K/2- I
(12.36)
An approximate expression for the second integral in the numerator of (12.19) may be found using similar arguments. When e = 0, f
sin # cos = r/2
2(l-cot2$cot2L)1/2sin&a(&)dd.
(12.37a)
•V/2- i
Approximating (l-cot 2 #cot 2 0 1 / 2 by a step function, this becomes f J
sin#cos£a(#)d#d£^ f J
A(ti\i)
TT/2-
2sin#a,(&)dd.
(12.37b)
i
Setting I = 0, the integral is //4(tilt)
2cose/r(l — cot 2 ^cot 2 e)
= /
sint?a(^)(i^,
(12.37c)
which is approximately sin # cos £ a(>&) d$d£ « /"Ir/2 2cos iff sin d a(d) rf*. (12.37d)
f
J
ir/2-
e
Combining both shadows and using the square-boundary approximation gives f
J
sin# cos £a(&)ddd£
A(ti\i)
« r J
TT/2-
2cos if/ sin '&a('&)d'&+ C* t
J
ir/2-
(l-cos
I
(12.37e) The first term on the right-hand side of (12.37e) is the same as the
342
12. Photometric effects of large-scale roughness
right side of (12.37d), which is an approximation to the right side of (12.37c). Similarly, the second term on the right of (12.37e) is the same as the right-hand side of (12.37b), except for the coefficient that accounts for the fact that only part of the illumination shadow contributes to the integral, and may be replaced by the right side of (12.37a). Thus, we obtain
f
J y4(tilt)
A
TT/2
/2-e
Ar/2- I
where the value of sin'Kl — cot 2 #cot 2 e) 1/2 lies between 7r/2 and TT, and that of sin~ 1 (l-cot 2 ^cot 2 0 1 / 2 lies between 0 and v/2. The next step is to carry out the integration over # in equations (12.35), (12.36), and (12.38). Let the average value of any function be defined as (12.39) We will obtain approximate analytic expressions that are valid for I and i near 90° by expanding the integrals in Taylor series in cot e and cot 6. This gives, for the integral in (12.35), •M(tilt)
-7r
-,
2
- —expV TT
/ \
2
^—.
TT
rr—cottt
\] \
.
, 1O . . .
(12.40)
y
'
12.B. Derivation
343
The derivation up to this point has not depended on 0 being small. This assumption will now be used for the first time. Using the slope distribution function (12.3) and applying (12.4) and (12.5), it is found that, to second order in 0,
(12.41a) (12.41b) (12.41c)
2
(12.41d)
Then (12.40) becomes
f
JJ A(tnt)
A(tnt)
cos#
X 2 - e x p l - —cot0cot e | - ^ e x p | -
— cot0cot i)\.
(12.42)
Expressions (12.36) and (12.38) may be evaluated in a similar way to give
f
J
a
A(ti\t)
^77 2 - e x p j
cot0cot e)- — expl
cot0cot i)\
(12.43)
and f
J
A(tili)
sin & cos £a(>&)ddd£
= [(TT tan 0 ) / ( l + 77 tan220) 0)V 2 cot 2 0cot 2 e
X cosii/rexp
L
\
IT
+ sin 2 y expl- ^ cot 2 0cot 2 i) I.
(12.44)
344
12. Photometric effects of large-scale roughness
Let x(d) =
= ex
P|
(12.45a) (12.45b)
cot 0 cot * I,
and (12.45c) Then the approximate analytic expression for (12.19) is cos I + sin I tan 0 (12.46) Using the identical procedure to evaluate (12.20) gives cose+sinetan0
£ 2K( e ) - s i n 2v( i / f / 2 ) £2V2 ( 0 ' ' * J, x (12.47)
If desired, expressions correct to higher order in 0 may be found, but such a complication probably would not be justified because of the assumption that interfacet scattering can be ignored. Equations (12.46) and (12.47) describe the effective tilt of the surface. Although they are mathematically complicated, their behavior is relatively simple. When I and e are smaller than about 7 7 / 2 - 0 , Moe ~ PoxW a n d Me — Px(9\ However, when either I or e exceeds about 7 7 / 2 - 0 , the effective surface tilts toward the source or detector by about 0, except that if both I and e are large and ^ is close to 77 the effective tilt angle goes to zero. Finally, from (12.46) and (12.47) the effective cosines at ij/ = 0, which appear in the shadow factor S, can be evaluated: 2
cose + sinetanfl
iy
(12.48)
and, by symmetry, cos i + sin I tan 6
Z
(12.49)
12.C. Applications to planetary photometry
345
Hence, from (12.28), S( i, 6, iff) =
firr
jfir —
(12.50) where (12.51) 12.B.3. The case when e <, / Similar reasoning when e < £ leads to the following expressions: \
cos I 4- sin £ tan 0
E2(i)-sin2(il//2)E2(e) ,..—. . ^ . , (12.52)
- cosil/E2(L) +sin2(il//2)E2(e)] (12.53)
(12.54) where iiOe(0)9 ne(0), and /((/r) are the same as for the i < e case and are given by (12.48), (12.49), and (12.51), respectively, except that /(0 is now to be interpreted as the fraction of the visibility shadow hidden in the illumination shadow. 12.C. Applications to planetary photometry 12.C.1. Disk-resolved photometry The equations derived in Section 12.B can be used to calculate the effects of macroscopic roughness on light scattered by a surface having an arbitrary diffuse-reflectance function. These results will now be applied to the approximate analytic bidirectional-reflectance equation (8.89). For a surface characterized by a mean roughness slope angle 0,
346
12. Photometric effects of large-scale roughness
this equation becomes
)
(12.55) without loss of generality, it may be assumed that g > 0. When I < e, or A < - g / 2 , juOe, /xe, and SU,e,g) are given by equations (12.45)-(12.51); when i > e, or A > - g/2, /x,Oe, jue, and SO, e,g) are given by equations (12.45) and (12.48)-(12.54); the mean slope angle 0 is defined in (12.5), and the quantities inside the curly brackets are defined in Chapter 8. Note that (12.55) satisfies the reciprocity requirement (Section 10.C), as may be verified by writing down the detailed expressions for the various terms in the reflectance. If this is done, it must be remembered that if I < e for a given set of angles, then the reciprocal configuration will have I > e. The approximate formalism developed in this chapter has been compared with detailed calculations of the brightness of an artificial rough surface generated by a computer model (Helfenstein, 1988), and the predictions of equation (12.55) have been found to be generally consistent with the model. The main deficiency of this model is its neglect of interfacet multiple scattering. This will be negligible if the reflectances of the facets are low, but it can be important if the surface is bright, especially at high phase angles where many shadows are visible but are partially filled in. If one is attempting to fit a theoretical curve to a bright surface, interfacet scattering will have the effect of making the reflectance of a surface higher at large phase angles, which will lead to a 0 that is too small. Hence, any deduced 0 should be regarded as a lower limit to the true value. Buratti and Veverka (1985) investigated the effects of shadowfilling. They used an exact theoretical model for shadowing in a crater of hemi-elliptical cross section, with a depth-to-diameter ratio of 0.4, and calculated the brightness assuming that the surface facets scattered light according to the Lommel-Seeliger law for a dark (r = 0.02) surface, and Lambert's law for a bright (r = 0.95) surface. This model neglected interfacet scattering.
12.C. Applications to planetary photometry
347
Buratti and Veverka then calculated the fractional changes in brightness that the crater caused. The changes are shown as the solid lines in Figure 12.4. They also measured the reflectances of real, highand low-albedo surfaces containing a similarly shaped crater. The fractional changes caused by the craters are shown as the dots in the figure. Theoretical and measured changes were similar for the dark surface, but differed by about 2% for the bright surface, which can be attributed to the effects of shadow-filling. Errors of 2% are well within the accuracy with which absolute reflectances can be measured in most situations. Hence, even for bright surfaces, a value of 6 deduced from photometry probably will not be seriously in error, especially when the slopes are not too large. In order to illustrate the effects of roughness on the reflectance, let us see how the distribution of brightness across the surface of a planet is altered by increasing 0. These changes are illustrated in Figures 12.5 and 12.6, which show relative reflectance profiles across the disks of hypothetical planets of varying roughnesses and albedos. Figure Figure 12.4. Comparison between predicted and measured changes caused in the reflectance by roughness for a bright (r — 0.95) and a dark (r = 0.02) surface. The dots are the measured fractional changes in percentage, and the lines are the predicted changes. From Buratti and Veverka (1985), courtesy of Academic Press, Inc.
c a
1
6
Dark Sample
-
TS
c .2
I«
y
2—
0 i
i
i
20
40 g (degrees)
i 60
348
12. Photometric effects of large-scale roughness Figure 12.5. Effect of macroscopic roughness on the brightness profiles of a low-albedo (w = 0.25) planet at three phase angles g. For simplicity, the surface is assumed to be covered with isotropically scattering particles, and the opposition effect is neglected. Solid lines, radiance factor for 6 = 25°; dashed lines, radiance factor for 8 - 0. (Top) Profiles along the luminance equator. (Bottom) Profiles along the central meridian of the illuminated part of the disk. From Hapke (1984), courtesy of Academic Press, Inc. 0.025
-90
-60
-30
0
30
60
90
Longitude (degrees)
0.015
0.01 O
c
B o o 0.005
30
60
Latitude (degrees)
90
12. C. Applications to planetary photometry
349
12.5 (top) shows the brightness as a function of longitude along an equatorial scan on a dark, uniform planet whose regolith is composed of particles of mean single-scattering albedo w = 0.25. Profiles for three phase angles are given for two values of 0. Figure 12.5 (bottom) shows the brightness as a function of latitude along the central meridian of the illuminated crescent of this planet for the same phase angles and slope angles. Figure 12.6 gives similar profiles for a bright Figure 12.6. Same as Figure 12.5 for a high-albedo (w = 0.95) planet. From Hapke (1984), courtesy of Academic Press, Inc.
-90
-60
-30
0
30
60
90
Longitude (degrees)
0.2
0.15 CD
o
cc 0.05
"
30
60
Latitude (degrees)
90
350
12. Photometric effects of large-scale roughness
planet with w = 0.95. For simplicity, the profiles were calculated from equation (12.55) assuming isotropic scatterers and neglecting the opposition effect. Note that under most conditions the effect of increasing roughness is to decrease the reflectance. This decrease is especially important at large phase angles, where shadows hide much of the surface. It also occurs at the limb of the low-albedo planet, where the surface elements are selectively tilted toward the observer, eliminating the limb spike caused by the Lommel-Seeliger law. However, near the limb and terminator of the high-albedo planet seen at small phase angles, the effective tilt causes an increase in brightness with increasing roughness. The brightness of a low-albedo planet is governed primarily by the Lommel-Seeliger law, which is independent of latitude (Section 8.G.2). Hence, a dark planet with a smooth surface would not be expected to Figure 12.7. Measured and predicted brightness distributions along the equator of Mercury at g = 77°. The line is the theoretical brightness for 6 = 20°; the crosses are data obtained by the Mariner 10 spacecraft. From Hapke (1984), courtesy of Academic Press, Inc.
-60
-30 Longitude (degrees)
12.C. Applications to planetary photometry
351
exhibit polar darkening. Roughening the surface causes the brightness to decrease with latitude. However, at small phase angles this decrease does not become pronounced until high latitudes, in agreement with observations by Minnaert (1961) for the moon. As the phase angle increases, the polar darkening moves to lower latitudes. The theoretical profiles are compared with several planetary data sets taken by spacecraft in Figures 12.7-12.9. Figure 12.7 shows an equatorial scan, and Figure 12.8 a meridional scan of Mercury using data from the Mariner 10 mission (Hapke, 1984). Figure 12.9 shows an equatorial scan of Europa (Domingue et al., 1991) based on Voyager measurements. These equations have also been applied to the analysis of planetary data in papers by Helfenstein (1986), Helfenstein and Veverka (1987), Veverka et al. (1988), McEwan (1991), and others. Figure 12.8. Measured and predicted brightness distributions along the A = -50° meridian of luminance latitude of Mercury at g = 77°. The line is the theoretical brightness for 0 = 20°; the crosses are data obtained by the Mariner 10 spacecraft. From Hapke (1984), courtesy of Academic Press, Inc.
.01 —
30
60
Latitude (degrees)
352
12. Photometric effects of large-scale roughness 12.C.2. Disk-integrated photometry
From equation (10.40) the integral phase function of a uniform, rough-surfaced, spherical planet of radius 31 observed at phase angle g is
ir/2-g
r/2
J
A=-TT/2 L
1
JrR(A,L,g)ti3?2cosLdLdA
=-ir/2
rR(A,L,g)cosAcos2LdLdA,
I
0
(12.56) where
/
/
r/2rR(A,L,0)cosAcos2LdLdA
(12.57)
A=-ir/2' L = 0
is the physical albedo. In these equations we have used the fact that Figure 12.9. Measured and predicted brightness distributions along the luminance equator of Europa at g = 10.5°. The line is the theoretical brightness in relative units for 6 = 10°; the crosses are data obtained by the Voyager spacecraft. From Domingue et al. (1991), courtesy of Academic Press, Inc. 1.0
I
-
0.6
-
I
.
1
1 '
1
1 •
1
1 •
1
1
i
I
.
I
.
0.8
1
0.2
I
0.4 —
A —
-
. -90
. 1 . -60
1
1 . -30
1
1
0
I
.
1 1 30
Luminance Longitude (degrees)
60
90
12. C. Applications to planetary photometry
353
the bidirectional reflectance must be symmetric with respect to the northern and southern hemispheres and reciprocal with respect to the A = - g/2 meridian. Equation (12.55) was inserted into (12.56) and (12.57) and integrated numerically for values of 0 up to 60°.It was found that to a good approximation the physical albedo can be written (12.58) and Q(g,w,0) = Jr(g,d)Q(g9w90),
(12.59)
where r0 is the diffusive reflectance, and 4>(g, w,0) is given by (10.47); in which Ap(w,0) is given by (10.38). The quantities W(w,0) and c^tg,0) are, respectively, the factors by which the physical albedo and integral phase function of a planet with a smooth surface may be corrected for effects of macroscopic roughness. These correction factors are plotted in Figures 12.10 and 12.11. It was found that to better than 1% the numerical values of %?(w,0) can be represented by the empirical expression ^(w,0) = 1 - (0.0480 + O.OO4102)ro - (0.330 - 0.004902)r2, (12.60) where 0 is in radians. The factor ^(g,d)
is tabulated in Table 12.1. It
Figure 12.10. Physical-albedo correction factor &(w,d) plotted against the single-scattering albedo for several values of 6. From Hapke (1984), courtesy of Academic Press, Inc.
0.9
U
0.8
0.7
0.6 0.2
0.4
0.6 W
0.8
354
12. Photometric effects of large-scale roughness Figure 12.11. Integral-photometric-function correction factor X(g, 6) plotted against phase angle for several values of 0. This quantity is the ratio of the normalized integral brightness of a planet with a rough surface to that of one with a smooth surface. From Hapke (1984), courtesy of Academic Press, Inc.
0.8
0.6
0.4
0.2
30
60
90 120 g (degrees)
150
180
Table 12.1. Integral-phase-function roughness correction factors X(g, 6) d (deg) g (deg)
0
10
20
30
40
50
60
0 2 5 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180
1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00
1.00 0.997 0.994 0.991 0.988 0.986 0.984 0.982 0.979 0.974 0.968 0.959 0.946 0.926 0.894 0.840 0.747 0.590 0.366 0.128 0
1.00 0.991 0.981 0.970 0.957 0.947 0.938 0.926 0.911 0.891 0.864 0.827 0.777 0.708 0.617 0.503 0.374 0.244 0.127 0.037 0
1.00 0.984 0.965 0.943 0.914 0.892 0.871 0.846 0.814 0.772 0.719 0.654 0.575 0.484 0.386 0.290 0.201 0.123 0.060 0.016 0
1.00 0.974 0.944 0.909 0.861 0.825 0.789 0.748 0.698 0.637 0.566 0.487 0.403 0.320 0.243 0.175 0.117 0.069 0.032 0.0085 0
1.00 0.961 0.918 0.866 0.797 0.744 0.692 0.635 0.570 0.499 0.423 0.346 0.273 0.208 0.153 0.107 0.070 0.040 0.018 0.0047 0
1.00 0.943 0.881 0.809 0.715 0.644 0.577 0.509 0.438 0.366 0.296 0.231 0.175 0.130 0.094 0.064 0.041 0.023 0.010 0.0026 0
12.C. Applications to planetary photometry
was found that for g<60°, empirical function exp j -0.320 tan0tan |
355
can be approximated by the -0.520tan0tan||;
(12.61) if g > 60°, this equation overestimates •#", and Table 12.1 should be used. Note that macroscopic roughness causes a small opposition effect in the integral phase function. However, this opposition effect is quite broad and cannot be separated from the rest of the phase curve. At small phase angles the regions that are strongly shadowed tend to be close to the limb, where their effects are minimized by the factor JX in the integrand for
150
356
12. Photometric effects of large-scale roughness
are oriented more directly toward the sun. The major effects of roughness are seen when g > 90°, when the heavily shadowed areas are near the center of the disk. At very large phase angles the brightness of a rough planet is only a few percent of that of a corresponding smooth planet, thus accounting for the observation, known since antiquity, that the moon is invisible when less than about 1 day from new. Fitting a theoretical integral phase function to pbservational data typically involves finding six parameters: w, 2?0, h, 0, and two or more parameters to describe pig). Fortunately, the various parameters affect different parts of the phase curve, so that they can be fairly well constrained. The single-scattering albedo w controls the height or absolute magnitude of the phase function and also of the physical albedo. The opposition-effect parameters BQ and h control the shape at small phase angles, and observations at g < 1° are needed to fix these quantities reliably. The mean slope angle 0 most strongly influences the amplitude of the phase curve at large phase angles, so that measurements of the integral brightness at g > 90° are needed to Figure 12.13. Measured and predicted integral phase functions of Europa. The line is the theoretical phase function for 6 = 10°; the crosses are combined telescopic and Voyager data. From Domingue et al. (1991), courtesy of Academic Press, Inc.
120 Phase Angle (degrees)
12.C. Applications to planetary photometry
357
constrain this parameter. The single-particle scattering function essentially fine-tunes the shape of the integral phase curve. Although it is possible to fit these equations to observations of a wide variety of bodies in the solar system, Domingue and Hapke (1989) have demonstrated that unique values of the parameters cannot be found from disk-integrated data alone. This is especially true if data are available over only a limited range of phase angles. However, if disk-resolved photometry is available, the parameters can be uniquely constrained. The theoretical integral phase function is compared with observations of Mercury, a low-albedo planet, in Figure 12.12 (Hapke, 1984) and Europa, a high-albedo body, in Figure 12.13 (Domingue et al., 1991). Equations (12.56)-(12.61) have been widely used in a number of analyses to describe the integral scattering properties of objects in the solar system, including studies by Hapke (1984), Buratti (1985), Helfenstein and Veverka (1987), Simonelli and Veverka (1987), Herbst et al. (1987), Veverka et al. (1988), Bowell et al. (1989), and Domingue et al. (1991).
13 Effects of thermal emission
13.A. Introduction
The region of the spectrum in the vicinity of 10 \x m wavelength is called the thermal infrared. It is important because many materials have strong vibrational absorption bands there (Chapter 3). In most remote-sensing measurements these bands can be detected only through their effects on the radiation that is thermally emitted by the planetary surface being studied. Many substances have overtone or combination bands at shorter wavelengths, and although the latter bands are observed in reflected light, their depths and shapes may be affected by the thermal radiation that is emitted by the material. Hence, even though the primary subject of this book is reflectance, it is important that the effects of thermal emission be discussed. It will be seen that most of the preceding discussions of reflectance also apply to emissivity at the same wavelength because of the complementary relation between the two quantities. Figure 13.1 shows the spectrum of sunlight reflected from a surface with a diffusive reflectance of 10%, compared with the spectrum of thermal emission from a black body in radiative equilibrium with the sunlight, at various distances from the sun. Clearly, thermal emission can be ignored at short wavelengths, and reflected sunlight at long, but at intermediate wavelengths in the mid-infrared the radiance received by a detector viewing the surface includes both sources. 358
359
13.A. Introduction
In this chapter, expressions will be derived for the radiant power received by a detector viewing a particulate medium, such as a powder in the laboratory or a planetary regolith, when either or both reflected sunlight and thermally emitted radiation are present. The reflectance models developed in previous chapters will be extended to include the effects of thermal radiation. However, the effects of large-scale surface roughness on thermal emission are beyond the scope of this book. A recent treatment of macroscopic roughness effects is that by Spencer (1990). There is a large body of literature that treats thermal emission and radiation transfer in planetary and stellar atmospheres (e.g., Chandrasekhar, 1960; Goody, 1964; Sobolev, 1975; Van de Hulst, 1980). Theoretical treatments of emittance from a particulate medium have been published by several authors, including Vincent and Hunt (1968), Conel (1969), Emslie and Aronson (1973), and Aronson et al. (1979). However, virtually all of the published theories of emittance by powders derive only the hemispherically integrated radiance, whereas it is the directional emittance that is measured remotely. The situation in which both reflectance and emittance contribute to the radiance has hardly been discussed at all. Figure 13.1. Comparison of sunlight reflected (solid lines) from a surface with a visual albedo of 0.1 with the radiation thermally emitted (dashed lines) from the surface with an IR emissivity of 1.00, for three different distances from the sun. 1000
10
0.1
10
100
360
13. Effects of thermal emission 13.B. Black-body thermal radiation
Suppose a hollow cavity is surrounded by material that is optically thick at all wavelengths and is heated to a uniform absolute temperature T (Figure 13.2). Then it is found that the spectral radiance in the cavity is given by (13.1) where £/(A,D is the Planck function, (13.2) hQ is Planck's constant (/io = 6.626 XlO 34 Jsec), c0 is the speed of light, and k0 is Boltzmann's constant (kQ = 1.381 XlO" 23 JK" 1 ). The quantities c1 = 2vh0cl = 3.742XlO"16 Wm2 and c2 = hQco/kQ = 0.01439 mK are called, respectively, the first and second radiation constants. The unit of £/(A,D is power per unit area per unit wavelength interval. The radiance in the cavity is found to be independent of direction, position, shape of the cavity, and composition of its walls. Figure 13.2. Thermal radiance inside a black-body enclosure whose walls are at temperature T and the power per unit area emerging from a small hole in the wall.
1 3.B. Black-body thermal radiation
361
At short wavelengths the exponential in the denominator of (13.2) becomes large, and U(X,T) - ^ ^ e - W A * o r (13 3a) A At long wavelengths this exponential may be expanded as ex ^ 1 + x, to give
tf(A,7>^£oI,
(13.3b)
which is known as the Rayleigh-Jeans law. By setting the derivative dU(\,T)/d\ = 0, it may be shown that the Planck function has a maximum at A = 2898 /T Aim, (13.3c) where T is degrees Kelvin. This relation, which is known as the Wien displacement law, shows that the wavelength of the maximum shifts toward shorter values as the temperature increases. If a small hole is drilled through one of the walls of the cavity, the radiant power per unit area emerging into the hemisphere above the hole is />em( A, T) = /
But x\ex
-I)'1
= x*e~x(l= x3e~x(l + e~x + e~2x = x3(e~x + e~2x + e~3x
362
13. Effects of thermal emission
Hence,
(>-l)-1&= £ foe'"*- E ^ - £ , which gives V(T)
=
(13.4)
where er0 = 2 T T 5 ^ / 1 5 ^ = 5.671 Xl(T 8 Wm" 2 K" 4 is the StefanBoltzmann constant. Equation (13.4) is the Stefan-Boltzmann law. The unit of V(T) is power per unit area. The corresponding integrated radiance inside a thermodynamic enclosure is I{T) = V(T)/TT. 13.C. The emissivity 13.C.1. The emissivity and emittance The power thermally emitted by a surface is often called the emittance. If the emittance of the surface of an optically thick sample of real material is measured, it is found that the spectrum generally is similar to the Planck function, but usually is smaller by an amount that may vary with wavelength. The ratio of the actual power f/a(A, T) to that emitted by an ideal black surface is the spectral emissivity:
If e is independent of wavelength, the surface is called a gray body. Similarly, the integrated emissivity e is the average emissivity weighted by the thermal spectrum:
f£
(13.6)
Just as there are several kinds of reflectances that are distinguished by the degree of collimation of the source and detector, several kinds of emissivities may be defined, depending on the degree of collimation of the detector. These are the directional, conical, and hemispherical emissivities, denoted, respectively, by ed, ec, and eh. Yet another kind of emissivity refers to the radiation emitted by a particle. In this book, s with no subscript will usually denote the emissivity of a single particle.
13.C The emissivity
363
13.C.2. The emissivity of a solid, smooth surface
It is found experimentally that the radiance inside a thermodynamic enclosure is independent of the nature of composition of the walls. Suppose the walls are smooth and polished. Then the radiant power incident per unit area on a wall inside such an enclosure is [/(A, T). A fraction Se(A) of this power is reflected, where Se(\) is the hemispherically integrated Fresnel reflection coefficient of the wall (Chapter 5), and 1 — Se(X) is absorbed. Thus, if the radiance inside the cavity is to be independent of the composition of the wall, the power absorbed must be exactly balanced by the power emitted by the wall. That is, for a specularly reflecting surface, eh(\)U(\,T) = [l- Se(\)]U(\,T\ so that 6*(A)-l-S f f (A). (13.7) Equation (13.7) is one form of a relationship known as Kirchhojfs law. This law will be discussed in more detail in Section 13.D. Although there are many kinds of emissivities, the one defined by equation (13.7) is usually the quantity that is discussed in most physics textbooks. 13.C.3. Emissivity and emissivity factor of a particle
Because we are interested in the thermal emission from particulate media, it is necessary to know the emissivity of a particle. This quantity can be found by considering the energy balance of a particle located inside a thermodynamic enclosure. As usual throughout this book, except where explicitly stated otherwise, it is assumed that we are dealing with ensembles of randomly oriented particles. As discussed in Chapter 6, the geometric cross sections and efficiencies of particles are taken to be equivalent to their rotationally averaged values, and so are independent of the direction of the incident radiance. Suppose an enclosure contains a particle with geometric crosssectional area
364
13. Effects of thermal emission
ance inside the enclosure is not to be altered by the particle, we must have = QAW(13.8) An alternate proof of this equality for homogeneous spheres has been given by Kattawar and Eisner (1970). Define the particle emissivity factor E(A) as E(A) =
^_L.
(13.9a)
Then
where w(A) is the particle single-scattering albedo. For large particles close together, QE = 1 and E = e. 13.C.4. The thermal source Junction
Suppose a small volume element Av contains a number of different types of particles. Then the power emitted per unit wavelength per unit solid angle from the volume is where N is the number of particles per unit volume, and the subscript / denotes the type of particle. It was shown in Section 7.2 that one of the quantities appearing in the radiative-transfer equation is the power thermally emitted per unit volume into unit solid angle. This quantity is known as the thermal volume emission coefficient and is
denoted by FT. Hence,
f
,
,
f
)
,
(13.10)
where K(X) is the volume absorption coefficient of a particulate medium [equation (7.37)]. Because of the assumption of random orientation, FT is independent of direction, although it may depend on position. Similarly, the volume thermal source function ^T = FT/E was defined in Chapter 7, where E is the volume extinction coefficient
13.G The emissivity
365
[equation (7.35)]. Equation (13.10) shows that for a collection of particles the volume thermal source function is K(k) U(\,T) *T-
E
TT
•
Define the volume emissivity factor of a particulate medium as
(13.11) where 5 is the volume scattering coefficient [equation (7.36)], w is the volume single-scattering albedo [equation (7.17)], and y is the volume albedo factor. If the particles are all of one type, are large compared with the wavelength, and are close together, so that QE = 1, then Thus, the thermal source function of a particulate medium can be written or
—
v / JT( \ T\ — J--±—-LTT(\
T\
(\*\
io\
When equations (13.10M13.12) are applied to a medium of large, closely packed particles, the modifications discussed in Chapter 7 apply-namely, diffraction is ignored, so that QE = 1, and N must be replaced by the effective particle density NE [equation (7.34)]. 13.C.5. The directional emissivity of a particulate medium
The directional emissivity sd(e,\) is the ratio of (1) the radiance /(e, A, T) emerging from the surface of a particulate at a uniform temperature T into a given direction making an with the zenith to (2) the thermal radiance emerging from body at the same temperature:
thermal medium angle e a black
In this section, the directional emissivity of an infinitely thick particulate medium of isotropic scatterers will be calculated using the method of embedded invariance. This method, which was used previously in Chapter 8 to calculate the bidirectional reflectance, is based on the principle that if a thin layer of particles is added to the top of
366
13. Effects of thermal emission Figure 13.3. Schematic diagram of the five first-order changes in the thermally emitted radiance caused by adding a thin layer of optical thickness AT to the top of an infinitely thick medium.
13.C. The emissivity
367
an infinitely thick medium consisting of the same type of particles, neither the emittance nor the reflectance will be changed. We will calculate the first-order changes caused by the addition of such a layer and then require that the sum of all these changes must equal zero. For convenience and economy of notation, the explicit dependences of the various quantities on A and T will be dropped. Figure 13.3. (cont)
Ax
Tk: J
:;••;:
r ( i , e)
.^——*^_ s
(c)
::ed(e'
(d)
. .^^»
368
13. Effects of thermal emission
Consider a semiinfinite medium consisting of N particles per unit volume, with bidirectional reflectance r(i,e,g). Because the particles emit and scatter isotropically, r(l,e,g) and ed(e) are independent of azimuth or phase angle and may be written r(/xo,A^) and ed(fi), respectively. Let a layer of thickness Az and optical thickness AT = NaQELz = EAz <: 1 be added to the top of the medium. Assume that AT is so small that interactions of light with the layer involving powers of AT greater than 1 can be ignored. Then the layer will cause five separate changes proportional to AT in the emitted radiation. These changes are shown schematically in Figure 13.3: (1) Radiance sd(e)(U/7r) emitted by the lower medium into a direction fte making an angle e with the vertical is attenuated by extinction in the added layer (Figure 13.3a). The radiance emerging ed(e)(U/7r)erAT/"~ed(e)(U/TTXIfrom the upper layer is I = ) , to first order in AT. Hence, the change due to this effect is U
AT
-
—
(13.14a)
(2) The added layer emits an additional amount of light into the direction ft6 (Figure 13.3b). Consider a cylindrical volume coaxial with the emitted ray with cross-sectional area a and length AZ/JJL, Figure 13.3. (cont)
13.C. The emissivity
369
where /JL = cos e. Then the radiance emitted toward fte by the particles in this volume is A/ 2 = NcrQF—8—
=
— U—.
(13.14b)
(3) The added layer emits an amount of light in the downward direction. This light is scattered by the lower medium into the direction fte (Figure 13.3c). Consider a cylinder of area a and length A z / / i 0 coaxial with a direction ft. making an angle I with the normal to the layer, where JU,0 = COS£. This cylinder emits a radiance (^'/TT)U(AT/> 0 ) into this direction toward the lower medium. The medium scatters a fraction rifi^fi) of this radiance into the direction ftc. The total contribution of this effect is the integral over all angles I:
W
%U^-r{^^)ddi.
(13.14c)
where dCli = 2TTsin idi = —lirdfiQ. emitted by the lower layer into a direc(4) Radiance ed(eXU/ir) tion ftc/ making an angle e with the vertical illuminates a cylinder of area a and length Az/fi in the added layer coaxial with the direction of this light is scattered fte (Figure 13.3d). A fraction (w/4TT)(AT/fi) by the particles in this cylinder toward fl e . The total contribution of this effect is the integral over all angles e: &
(1314d)
where d£lj = 2TTsin ede=—2Trd\x and /J! = cos e'. (5) Radiance sd(e')(U/IT) emitted by the lower layer into a direction ft6> making an angle e! with the vertical illuminates a cylinder of in the added layer coaxial with the area or and length Az/fiQ direction H. (Figure 13.3e). A fraction (w/47r)(Ar/)Lt 0) of this light is scattered by the particles in this cylinder parallel to ft; toward the lower layer. A fraction r(/i o,)it) is scattered by the lower medium into the direction ft6. The total contribution of this effect is the double integral over all angles e and I: ^
(13.14c)
370
13. Effects of thermal emission
The sum of all the changes A/x through A/5 must be zero. Hence, after dividing through by [/AT/TTJU,, we obtain
(13.15) Now, it was shown in Section 8.G.3.b that a medium of isotropically scattering particles has a bidirectional reflectance
where
//(JLI)
satisfies the integral equation
Hence,
and (13.14) becomes
]
(13.16)
Integrating (13.16) with respect to /x,,
f\M
d/x = [/^if (n) d/t] [«r + y
and solving for the integral of the emissivity gives l l dp = %( J H{ix) dy\\ -L \J \ H(n) dJ o o L o J But according to (8.32), Jj\d(li)
.
(13.17)
13. C. The emissivity
371
Inserting this into (13.17) gives %
2
(13.18)
Substituting this result into (13.16) gives (13.19) Note that, like equation (8.51) for the reflectance, which was derived in the same manner, (13.19) is an exact, general solution for ed and makes no assumptions about the medium, other than that it is composed of particles that emit and scatter isotropically. If approximation (8.55) is used for H(fi), (13.19) becomes (13.20) Thus, the radiance emerging from the surface of an optically thick particulate medium at uniform temperature T is f).
(13.21)
The directional emissivity ed is plotted versus e for several values of w in Figure 13.4. When w is small, ed = 1 at all angles, and the surface Figure 13.4. Directional emissivity of a particulate medium as a function of the angle of emergence for several values of the single-scattering albedo.
0.8
0.6 w«0.84 0.4
w=0.99
0.2
30
60 e (degrees)
90
13. Effects of thermal emission
372
emits like a black body. Hence, for low-albedo materials, the assumption that the emissivity is independent of angle is a good approximation. As w increases, ed decreases, and also ed decreases as e increases. The reason for the dependence on e is that the multiply scattered flux is important when the albedo is large. This flux decreases toward the surface because of leakage from the surface. As e increases, the field of view includes a greater contribution from the smaller flux closer to the surface. Very few measurements of the directional (as opposed to hemispherical) emissivity have been published. However, Jakowsky, Finiol, and Henderson (1990) have measured the directional emissivities of natural sand and playa surfaces, and they seem to be in qualitative agreement with the theoretical model of this chapter.
13.C.6. The hemispherical emissivity of a particulate medium
The hemispherical emissivity eh(\) of a particulate medium can be found by integrating the upward component of the emitted radiance, equation (13.21), over the upward hemisphere. The power emitted per Figure 13.5. Hemispherical emissivity of a particulate medium versus the single-scattering albedo.
0.8
0.6
0.4
0.2
I
0.2
.
I
0.4
0.6 W
0.8
13.D. Kirchhoffs law
373
unit area is
(13.22) where Hx = f^Hi^ixd^ is the first moment of the H function. Hence, the hemispherical emissivity is (13.23) Using approximation (8.58b), H1-[l/(l + y)1Kl + ^rQ\ where ro = (1 — y)/(l + y) is the diffusive reflectance, gives (13-24) The hemispherical emissivity sh is plotted versus w in Figure 13.5. Note that when w and r0 are small, 8
-
13.D. Kirchhoff's law KirchhofFs law is an extremely powerful and useful rule which states that there is a complementary relationship between emissivity and reflectance. KirchhofFs law allows the emissivity to be calculated from the reflectance, which often is more convenient to measure in the laboratory. However, as we have seen, there are several different kinds of reflectances and emissivities, and it is not always obvious which ones form the complementary pairs. From the derivations of this chapter, the following quantities obey KirchhofFs law. In Section 13.C.1 it was seen that the hemispherical emissivity of a smooth surface is eh = l-Se,
(13.25a)
where Se is the integral of the Fresnel reflection coefficients over a hemisphere. From KirchhofFs law it follows immediately that the directional emissivity of a smooth surface is ed(e) = l-R(e)9
(13.25b)
374
13. Effects of thermal emission
where R{ e) is the average of the Fresnel reflection coefficients over the two directions of polarization. From equations (13.8) and (5.2), the emissivity of a single particle is £ = QA = QE-QS> (13.26) where QE and Qs are the particle extinction and scattering efficiencies, respectively. From equations (13.11) and (7.17), the volume emissivity factor of a particulate medium is (13.27) ^ = 1-H>, where w is the volume single-scattering albedo in the medium. From equations (13.19) and (10.19), the directional emissivity of a particulate medium of isotropic scatterers is
ed(e) = l-rhd(*)> ( 13 - 28 ) where rhd(e) is the hemispherical-directional reflectance of the medium. Hence, the directional emissivity may be calculated from a measurement of the hemispherical-directional reflectance. However, because rhd(e) has the same functional dependence on e as the directional-hemispherical reflectance rh(l) has on I, the latter quantity may also be used to calculate ed(e). From (13.23) and (10.22), the hemispherical emissivity of a particulate medium of isotropic scatterers is where rs is the spherical or bihemispherical reflectance. Hence, if it is desired to calculate the hemispherical emissivity from reflectance, the spherical or Bond albedo is the quantity that must be measured. As discussed in Chapter 11, this is inconvenient to do in the laboratory. However, according to (10.23), an approximate expression for the spherical reflectance is rs « r o{l - \[y / ( I + y)]}. Substituting this into (13.29) gives the following approximate expression for the hemispherical emissivity:
(
§)
(13.30)
If w and r0 are small, For media in which the scatterers are approximately isotropic, several different kinds of reflectances reduce to the diffusive re-
13.E. Combined reflection and emission
375
flectance at certain angles [equation (10.24)]. Hence, the spherical reflectance and hemispherical emissivity may also be calculated from a measurement of one of those reflectances at the appropriate angle. 13.E. Combined reflection and emission from a medium with a temperature gradient When a planetary surface is observed in the mid-infrared, the detected radiance may be a mixture of reflected sunlight and thermal emission. In this section we will calculate the radiance emerging from the surface of a plane-parallel, semiinfinite medium illuminated by a collimated source, and including thermal emission. The radiance in the medium is governed by the equation of radiative transfer, equation (7.21), with &T given by (13.12),
^
£),
(13.31)
where II is a direction making an angle # with the normal to the planes of symmetry. All quantities in (13.31) may be functions of A, but for economy of notation this explicit dependence is omitted. Suppose that a temperature gradient exists within the medium, producing a corresponding gradient in the thermal emission function. Equation (13.31) will be solved under the assumption that U(T,T) can be described to an adequate approximation by U(T,r) = U0(TQ) + Ux{T0)e-*'r,
(13.32)
where the quantity y is a dimensionless distance in units of the extinction length 1/E that characterizes the distance over which the temperature and the thermal-emission function change appreciably. For example, suppose that near the surface of the medium,
so that . dU dT
376
13. Effects of thermal emission
But, from (13.32), U{T,T)
= Uo + Uie~^
= UQ + Ut(l
-
Hence, Uo + Ux - U(T0)
and
u^-
The method of solution for this problem will be the same as used in Section 8.G, where the approximate analytic solution for the bidirectional reflectance was found. A first-order solution will be obtained with the two-stream method, assuming that the particles of the medium scatter isotropically. Using the method of successive approximations (Section 7.F.7), this solution will be substituted back into the radiative-transfer equation to obtain an improved solution, from which the emergent radiance will be found. Applying the multistream method (Section 7.F.5), in which the number of streams Jf=2, equation (13.31) with p(g) = 1 becomes
where Ix is the average radiance traveling into the upward hemisphere, and I2 is that traveling into the downward direction. As usual, let Ix = cp + A^> and I2 =
= - ?V + J ^ e ' T / l i 0 + h 2 ^
(13.33a) + u r/r
*- )>
(13-33b)
with the boundary conditions that
(13.34)
The solution to equations (13.33) that satisfies the boundary conditions is
+ j/ 2 e" T / / A °+ J / 3 + j / 4 e " T / ^ ,
(13.35)
13.E. Combined reflection and emission
377
where 1 77
w
An improved solution is next obtained by the method of successive approximations. The integral form of the equation of radiative transfer, equation (7.56), contains two terms, the multiple-scattering term / 4w J(T,ft')p(T',ft',ft)dft' and the source term ^ ( T , H ) . AS in Chapter 8, it will be assumed that most of the effects of nonisotropic scattering are due to the irradiance part of the source term. The two-stream solution for the integral of the intensity forisotropic scatterers is 4TT
^
yH{r)H{tl)Hi,
(1336)
where H(x) is given by the two-stream approximation for the H functions, H(x) = (1 + 2x)/(l + 2yjt). The details of the derivation of (13.35) and (13.36) are left as an exercise for the reader. The coefficient of / on the right-hand side of (13.36) is the analytic approximation for the bidirectional reflectance that was derived in
378
13. Effects of thermal emission
Chapter 8. This solution may be further improved by multiplying the particle phase function pig) by the expression describing the opposition effect, equation (8.90), to give
where B(g) = 5 0 / [ l + (l//i)tan(g/2)], and the physical interpretation of the parameters Bo and h is discussed in Section 8.H. Recall that the opposition effect occurs because of the overlapping of the extinction cylinders associated with the incident and emergent rays. Because there is no incident ray for thermal emission, an opposition effect can occur only in the reflected component, not in the thermally emitted component. Comparing (13.37) with equation (13.19), the expression multiplying C/Q/TT in the second term on the right-hand side of (13.37) is seen to be the directional emissivity associated with the constant component of the thermal emission. The effect of the temperature gradient is described by the third term on the right. Note that the component of the emissivity associated with the gradient has a different functional form than (13.19). If y
o
For isotropic scatterers, the integration is easily carried out with the aid of the defining equation for the H functions, equation (8.50). The result is (13.38)
13.F. Emittance spectroscopy
379
or, using the two-stream approximation to the H functions,
Pemv ~ y0 + yl ^+ -yf \1 + 6ir 0°W/o °+ 1T+ y^ m(i) / « / A ^ °t l0 1+l 2yix (13.39) The coefficient of Jfi0 in the first term is the directional-hemispherical reflectance rdh, and the coefficient of Uo in the second term is the hemispherical emissivity eh of a medium at constant temperature. All parameters in (13.37) and (13.38) are functions of A. If the equation of radiative transfer (13.31) is integrated over all wavelengths, equations almost identical with (13.37) and (13.38) are obtained, except that the parameters w, y, and p(g) are replaced by their wavelength-averaged values vv, y, and p(g), respectively. Thus, if V(T) can be represented by expressions of the form then the emergent wavelength-integrated radiance and hemispherical power are, respectively,
^
j
,
v
)
%
-
(13.40)
and (13.41) 13.F. Emittance spectroscopy
In applications of emittance spectroscopy to remote sensing, the radiance in the thermal infrared part of the spectrum emitted from the surface of a planet into a given direction is measured. A Planck function is then fitted to the overall data and is assumed to be representative of the temperature of the surface. Dividing the measured spectrum by the fitted Planck function gives the directional emissivity spectrum. An alternative procedure is to use the measured radiance to calculate a quantity called the brightness temperature. The brightness temperature is a useful concept for characterizing the radiance coming
380
13. Effects of thermal emission
from an object. It is defined as the temperature that a perfect black body of the same size and distance would have to have in order to emit the measured radiance. It is not necessarily related to the actual temperature of the surface. Thus, if /(A) is the radiance emerging from a surface, then the brightness temperature Tb is given by
which can be solved for Tb to give h c
In 1+
l ^(A)
If /(A) = erfC/(A,D/7r, the spectrum of Tb(\) is similar to that of ed(\), although the relationship obviously is not linear. The advantage of using the brightness temperature is that this procedure avoids possible errors caused by the subjective judgments involved in fitting the Planck function to the observed radiance. After the spectrum of either ed(\) or Tb(\) is obtained, it is then inspected for absorption bands that may be diagnostic of composition. An excellent discussion of the interpretation of emissivity spectra may be found in the work of Salisbury (1993). Because of the complementary relation between the emissivity and reflectance, exactly the same principles, as discussed in Chapter 10, apply to both. In particular, both the reflectance and the emissivity of a particulate medium are primarily controlled by the single-scattering albedo w (or, equivalently, the albedo parameter y), which in turn depends on the complex refractive index m = n + ik, the effective particle size
13.F. Emittance spectroscopy
381
emissivity is high, but the band has small contrast. A strong band whose center extends into the strong surface-scattering region is expressed as two peaks on either side of a dip (Figure 11.7c). In general, the contrasts in absorption bands observed in emissivity spectra are much smaller than the contrasts observed in reflectance, particularly in specular reflection from a polished surface. This point has been strongly emphasized by Conel (1969). Thus, there is an unfortunate paradox that often limits the usefulness of emittance spectroscopy for compositional remote sensing. In spite of the fact that the compositionally diagnostic restrahlen absorption bands are strong in the thermal infrared, the emissivity contrast that can be observed in a particulate medium may be very small. There are several reasons for the loss of contrast. First, as the particle size in a medium decreases, the contrast in the vicinity of the very strong restrahlen bands may decrease also (Hunt and Vincent, 1968; Salisbury, 1993) as the bands move from the strong into the weak surface-scattering region. If grinding moves the band all the way into the volume-scattering region, a restrahlen band may be expressed as a dip in large particles and as a peak in fine ones (Arnold and Wagner, 1988). If the medium consists of a mixture of coarse and fine particles, such effects may cause the bands to be almost unobservable. Further complicating this discussion is the fact that many media have particles that are smaller than the wavelength in the thermal infrared. The particle efficiencies and single-scattering albedos are reasonably well understood when their sizes are large compared with A. As was discussed in Chapter 7, when (D) «: A, it is not clear what effective particle size and refractive index should be used for the calculation of w in the radiative-transfer equation. However, the spectra seem to behave qualitatively as predicted from the assumption that the scattering from each particle is incoherent and quasiindependent of that from other particles. As seen in Chapter 5, absorbing particles smaller than the wavelength are almost perfect absorbers, which is equivalent to being almost perfect emitters, independent of wavelength. Second, as the particle emissivity decreases, the particle albedo increases, thus increasing the multiple scattering within the medium. It is sometimes stated in the literature that the loss of contrast is caused by void spaces in the medium acting like thermodynamic
382
13. Effects of thermal emission
enclosures. However, this is a misleading physical picture. The loss of contrast occurs because the increased multiple scattering partly cancels the decreased emissivity. Third, the loss of contrast in the restrahlen-band region can be seen from the mathematical form of the dependence of emissivity on single-scattering albedo. For example, consider eh. Figure 13.5 shows that eh is a monotonic, nonlinear function of w. When w is small, eh is large, and vice versa. The slope of the curve is small when w is small, and large when w is close to 1. Thus, a high spectral contrast is observed when w ~ 1, but the contrast is much smaller when w
1
=
2
Thus, AeH
Aw
If w
383
13.F. Emittance spectroscopy Figure 13.6. Change in the contrast of an absorption band as the band center moves from a wavelength where the emitted radiance is dominated by reflectance to domination by thermal emission. 30
1
-
20
o
1 0
\
-1 0 10
Figure 13.7. Spectral emissivities of polished quartz and quartz powders of different sizes and packings. The arrows show the locations of the Christiansen wavelength and the transparency feature (denoted by A c and A r , respectively). After Conel (1969); copyright by the American Geophysical Union. A
C 8
1400
T
Wavelength (jim) 10
1200
1000 Wave Number (crrr1)
12
800
14
600
384
13. Effects of thermal emission
the band increases, thermal emission fills in the band and decreases the contrast, until the band becomes unobservable at 5.8 /im. For wavelengths longer than about 6 /im the contrast is negative; that is, the radiance is dominated by thermal emission, and the band manifests itself as a weak maximum rather than a strong minimum. 13.G. Christiansen wavelengths and transparency features
Logan et al. (1973) and Salisbury and Walter (1989) have emphasized that the wavelength Ac corresponding to the Christiansen frequency (Section 4.E) of a substance is diagnostic of composition, and they have emphasized the potential of this technique for remote sensing in the thermal infrared. At the Christiansen wavelength of a particle, the real part n of its refractive index is the same as the average of that of the medium in which it is embedded. Because most scattering is produced by interactions of the light wave with discontinuities in n, Qs and w become very small near the Christiansen wavelength. This has two effects. First, r(A c )«:l and e ( A c ) - l . The Christiansen wavelength in quartz is shown in Figure 13.7. Second, because of the Figure 13.8. Correlation between the Christiansen wavelength and the transparency feature for silicates of different compositions. From Salisbury and Walter (1989); copyright by the American Geophysical Union. 1
13
1
1
1
1
1
+ +
E
• • •
••
•
12
—
X X X
X
x
X
•
X
-
X
0 Felsic rocks
"~
X Intermediate -
•
Mafic rocks
+ Ultramafic
11
1
7.5
1
1
l
i
8.0
l
t
.
i
1
i
8.5
Christiansen Wavelength (|im)
i
i
i
9.0
13.G. Christiansen wavelengths and transparency features
385
lack of scattering, the extinction length can be very long, so that much of the emerging radiation comes from deep within the medium. If the temperature increases with depth in the medium, as is often the case, a pronounced maximum in the emissivity spectrum will occur at the Christiansen wavelength. Silicates have very strong restrahlen bands between 8.5 and 12.0 jitm, depending on composition, because of the stretching of Si — O bonds. They also have slightly less intense bands between 16.5 and 25 /xm associated with Si — O — Si bending vibrations. In between the two bands the particles are in the volume-scattering region, producing a maximum in reflectance or a minimum in emissivity. Salisbury and Walter (1989) and Walter and Salisbury (1989) have pointed out that the wavelength of this peak or dip is diagnostic of composition, and they have termed it the transparency feature. The transparency feature in quartz is illustrated in Figure 13.7. Thus, the Christiansen wavelength and the transparency feature are correlated with each other and are diagnostic of composition. Figure 13.8 shows this correlation with composition for silicate rocks.
14 Polarization
14.A. Introduction
In the equations for the reflectance and emissivity of a participate medium developed in Chapters 7-12 it has been assumed that polarization can be neglected. For irregular particles that are large compared with the wavelength of the observation, this assumption is justified on the grounds that the light scattered by such particles is only weakly polarized. However, the polarization of the light scattered by a medium does contain information about the medium and thus is a useful tool for remote sensing. One of the advantages of using polarization is that it does not require absolute calibration of the detector, but only a measurement of the ratio of two radiances. The discovery that sunlight scattered from a planetary regolith was polarized was made as early as 1811 by Arago, who noticed that moonlight was partially linearly polarized and that the dark lunar maria were more strongly polarized than the lighter highlands. Subsequent observations of planetary polarization were made by several persons, including Lord Rosse in Ireland. However, the quantitative measurement of polarization from bodies of the solar system was placed on a firm foundation in the 1920s by the classical studies of Lyot (1929). This work was later continued by Dollfus (1956) and his colleagues. For a more detailed historical account, the reviews by Dollfus (1961, 1962) and Gehrels (1974) should be consulted. Other important contributions have been made by Gehrels and his co-workers 386
14.B. The Stokes vector
387
(Gehrels and Teska, 1963; Gehrels et al., 1964), who emphasized the variation of polarization with wavelength. Egan (1985) has discussed applications of polarization to terrestrial remote sensing. One of the triumphs of photopolarization in planetary remote sensing occurred when Hansen and Arking (1971) were able to account for the observed variation of the polarization of Venus as a function of wavelength and phase angle by a cloud of spherical particles of refractive index n = 1.44 and radius 1.1 fim. This was the key observation that led to the identification of the composition of the clouds as sulfuric acid (Young, 1973).
14.B. The Stokes vector
As we saw in Chapter 2, electromagnetic radiation consists of transverse waves of electric and magnetic fields vibrating perpendicularly to the direction of propagation. Because the magnetic vector is always perpendicular and proportional to the electric vector, the wave may be described by the amplitude and direction of the electric vector. The electric vector may be resolved into two perpendicular components, and in order to describe the wave completely the phase between the two components must be specified. If the phase between the two components of the electric vector is random, the wave is unpolarized. If the two components are exactly in phase, then they may be added to produce a resultant that always lies in one direction, and the wave is linearly polarized. If the phases of the two components differ by exactly 90°, the resultant may be described by a vector whose direction rotates and whose tip traces out an ellipse; in this case the wave is elliptically polarized. If the two components are of equal amplitudes and are out of phase by 90°, the resultant traces out a circle, and the wave is circularly polarized. Circularly polarized waves are further described as right-handed or left-handed. Unfortunately, there are two conflicting systems for describing the sense of handedness, or helicity. One system, which is commonly used by scientists and engineers working at radio frequencies, describes a circularly polarized wave as right-handed if a screw with a right-handed thread would advance in the same direction as the direction of propagation of the wave if turned in the same sense as the rotation of the electric vector. The other convention is usually fol-
388
14. Polarization
lowed by researchers working at shorter wavelengths in the ultraviolet, visible, or infrared, and it calls the sense of polarization right-handed if the screw would advance in the direction from which the wave is
coming. The convention being used should always be ascertained when consulting a reference. The complete specification of the intensity and state of polarization of light requires four quantities: the total intensity, the amount and direction of linear polarization, and the amount and helicity of circular polarization. (Although five quantities are listed, it may be shown that only four are independent.) These quantities may be combined into a four-dimensional vector called the Stokes vector. Suppose the electric field of an electromagnetic wave is given by two components, ,;(27ri//-27rz/A+5 JC )
and J(2>rrvt-2Try/\+8 )
parallel to two arbitrary axes, x and y, that are perpendicular to each other and to the direction of propagation, taken to be the z axis. Then the Stokes vector is given by
1=
2Ex0Ey0cos(8y-8x
(14.1)
2ExQEyOsin(8y-8x When a medium is illuminated by a partially polarized wave, both the incident and scattered radiances are Stokes four-vectors, and the reflectance is a 4 x 4 matrix called the Mueller matrix. Fortunately, this degree of complexity often is not needed in the interpretation of remote-sensing observations. In most laboratory and field measurements the illumination is sunlight, which is essentially unpolarized. However, the light scattered from the atmosphere is partially linearly polarized (e.g., Coulson, 1971), and that may need to be taken into account in measurements of the reflectance of the surface of the earth that are sensitive to polarization. Another impor-
14.C. Linear polarization of paniculate media
389
tant application in which the incident radiance is polarized is radar. The transmitted pulse is usually completely polarized, and the difference between the fractions of the power returned with the same and opposite senses of polarization as transmitted gives additional information about the medium (Evans and Hagfors, 1968). A general discussion of polarized reflectance and the Mueller matrix is beyond the scope of this book. This chapter will concentrate on the linear polarization in the radiance at optical wavelengths scattered by a medium of irregular particles large compared with the wavelength when illuminated by an unpolarized incident irradiance. Circular polarization will be ignored, because the few measurements that have been made have indicated that such polarization in the sunlight scattered from bodies of the solar system is extremely small, on the order of 10" 4 -10" 5 (Kemp, 1974). Unfortunately, the discussion will necessarily be limited, because even the apparently simple linear case is still poorly understood theoretically 60 years after Lyot's pioneering investigations.
14.C. Linear polarization of participate media
It is found observationally that when the incident light is unpolarized, the radiance scattered from a particulate medium is partially linearly polarized, and the plane of the polarized component is either parallel or perpendicular to the scattering plane. This allows the polarization, which is a vector quantity, to be represented by a scalar. Let I± (I, e, g) be the component of intensity perpendicular to the scattering plane, and /||U, e,g) the parallel component. Then the polarization is defined as
Thus, if the radiance scattered with the electric vector perpendicular to the scattering plane is larger than the radiance parallel, the polarization is said to be positive; if the opposite is true, the polarization is negative. With this choice of sign of equation (14.2), light specularly reflected from a plane surface (Fresnel reflection), and also light scattered by a small particle (Rayleigh scattering), will always
390
14. Polarization
have positive polarization. If the x axis of equation (14.1) is perpendicular to the scattering plane, and the y axis is in this plane, then P is the ratio of the second to the first component of the Stokes vector. The variation of polarization with phase angle of a particulate medium is typified by the curves of Lyot (1929) for the moon, shown in Figure 14.1. The curves for most of the airless bodies of the solar system are similar. Although these particular figures are for the integrated light from the entire body, the curves for individual areas are virtually identical (Dollfus, 1956). At zero phase angle the polarization is zero. The polarization is negative for small phase angles and goes through a minimum around P ^ — 1% at g ^ 10°.The polarization increases to zero at phase angles between about 15° and 30°, where the plane of polarization abruptly rotates and P becomes positive. It then goes through a maximum whose amplitude depends on the material, but is typically P ^ 10% at g around 100°, after which P decreases to zero again at large phase angles. The parameters describing the polarization phase curve are the amount Px and phase angle gx at which the maximum occurs, the amount Pn and phase angle gn at which the minimum occurs, Figure 14.1. Polarization versus phase angle for the integrated light from the moon. The crosses give the polarization before full moon, and the circles after (Lyot, 1929). 0.1
1
1
1
0
0.08
C
0.06 -
Or
0.04
9
'6 ' ' ' 0
0 •
0 0
f
+
9
0.02
-
- before full moon 0
•
-
+0
+
- after full moon
-
/
-0.02 30
60
90 120 g (degrees)
150
180
14.C. Linear polarization of paniculate media
391
and the inversion phase angle gt and the slope at ht = {dP/dg\g^ at the inversion angle. All of these parameters depend on particle size, composition, albedo, and porosity in a nontrivial way. If the polarization of light scattered from the surface of a nonopaque solid or liquid, such as a mineral or the surface of water, is examined, it is found that when the source and detector are in the vicinity of the specular configuration the polarization is large and positive, as predicted by the Fresnel reflection coefficients. However, away from the specular region the electric vector lies in the plane formed by the surface normal and the emergent ray, and the polarization is not zero at g = 0 (Figure 14.2). This residual negative polarization is caused by the light that has been volume-scattered within the material and is polarized by refraction as it leaves the surface, as described by the Fresnel transmission coefficients. However, the light scattered from a particulate medium has its plane of polarization parallel or perpendicular to the scattering plane, rather than with respect to the plane containing the exit ray and the surface normal. Furthermore, for low-albedo surfaces, such as lunar regolith, the polarization is a function only of phase angle and is independent of I and e, although in materials with higher albedos the polarization phase curve is somewhat dependent on I and e also. These observations suggest that the phenomena contributing to P(g) Figure 14.2. Polarization curve of light scattered from a glass plate viewed at e = 60°. 0.3
0.2
0.1
-
• -0.1
-0.2
-
" •
0
10
20
30
40 g (degrees)
50
60
70
80
392
14. Polarization
are mainly associated with single scattering by individual particles of the medium. In order to try to understand the polarization phase curve, at least qualitatively, the positive and negative branches of the curve will be discussed separately in the following sections. 14.D. The positive branch of polarization 14.D.1. Factors affecting the positive polarization
We will begin the discussion of the positive branch of the polarization phase curve by assuming that it is controlled primarily by the polarizing properties of the particles of the medium acting singly. Recall that the light scattered from a particulate medium is the result of four phenomena: (1) Light passing near the particle is diffracted. However, it was argued in Chapter 7 that diffraction may be neglected in a medium where the particles are close together. (2) Light that has been specularly reflected from the surface of the particle is positively polarized, in accordance with the Fresnel reflectance expressions (Chapter 4). (3) Light that has penetrated into the interior of a particle is negatively polarized by refraction as it leaves. In a regular particle, such as a perfect sphere, the plane of scattering is preserved, so that the transmitted light is strongly negatively polarized. If the particle is not opaque, the once-transmitted light is important for phase angles larger than 2# c , where # c is the critical angle for total internal reflection. Most authors assume that in an irregular particle the orientation of the scattering plane of the refracted light is randomized so that the volume-scattered light can be taken to be unpolarized in the first approximation. It will be shown later that this is not a valid assumption and that, in fact, the transmitted light controls Px and gx in nonopaque particles. (4) The final contribution is by light that has been multiply scattered by many particles. It will be assumed, as a first approximation in this section, that the planes of scattering between particles have been randomly rotated, so that this component is unpolarized. However, it will be seen in the discussion of the negative branch of polarization that this assumption probably is not valid for doubly scattered light.
14.D. The positive branch ofpolarization
393
If only the polarization by single-particle scattering contributes to the positive branch, then a theoretical expression for P may be derived as follows. In order that the expression may be applied to planetary regoliths, as well as laboratory surfaces, the effects of macroscopic roughness will be included. It has been shown that the radiance scattered by a medium is given, to a good approximation, by
xS(i,e,g),
(14.3)
where the quantities in this equation are as defined in Chapters 8 and 12. In equation (14.3) the radiance that has been scattered only once by an average particle of the regolith is described by wp(g). Let the portion of this radiance scattered with its electric vector perpendicular to the scattering plane of the medium be [wp(g)]±, and let that parallel to the scattering plane be [>vp(g)],|. Then if the incident radiance is unpolarized, and if the assumption that the multiply scattered radiance has no net polarization is valid, the two components of the scattered radiance are
+ w[H(» Oe)H(nLe)-l]}S(i,e,g)
(14.4a)
+ w[H(nOe)H(ne)-l])S(i,e,g).
(14.4b)
and
Hence, the polarization is P(i,e,g)
(14.5) which may be cleared of common factors in the numerator and
394
14. Polarization
denominator to give
(14.6) where A[Mp(g)] = [**(*)] x " [Mp(g)] „.
(14.7)
Note that P is independent of the shadowing function S(£, e,g). Thus, macroscopic roughness affects the polarization only through the effective tilt angles in the multiple-scattering contribution. Furthermore, because the opposition effect is negligible for phase angles larger than about 20°, to a good approximation in the positive branch,
If w is small, then the H functions are not very different from 1, and the polarization function is approximately
which is independent of I and e, as is observed for the moon (Dollfus, 1961, 1962). It is also independent of surface roughness. Thus, for dark surfaces the polarization of the light scattered by the medium is approximately equal to that of an average single particle of the medium. Similarly, from equations (12.56)-( 12.59), if the integral light from a body is being observed, the two components of its integral phase function are
x[l-sin|cos|ln(cot|)l
(14.10a)
14.D. The positive branch of polarization
395
and
X [l - sin | cos | ln(cot | ) ]
j
(14.10b) where £? is the radius of the body. Thus, because B(g)=^0 in the positive branch, w[p(g)-l]+4ro(l-ro) 16
sing+(-n--g)cosg
2 1
'
^
4-
(14.11)
"
Note that to the extent that the approximations of Chapter 12 are valid, this expression is independent of macroscopic roughness. If w <^c 1, terms of order rQ2 can be ignored, ro — w/4, and (14.11) reduces to (14.9), the polarization of an average regolith particle. For applications of polarization to remote sensing we wish to know what properties of the particles or the medium control Px and gx. Now, the scattering by an individual particle consists of two parts: Fresnel reflection from the surface, and light that has been refracted and scattered from the interior. Many theoretical models of polarization assume that the refracted light is randomly polarized and does not contribute to A[wp(g)]. Because gx occurs in the vicinity of 2&B, where &B is the Brewster angle, it is often assumed that in the positive branch of polarization, A[wp(g)] is controlled only by Fresnel reflection from the surfaces of the particles and that Px occurs at 2ftB. Therefore, let us investigate these assumptions quantitatively. We saw in Chapter 6 that the light transmitted through irregular, rough-surfaced particles retains some negative polarization. However, assume for the moment that the volume-scattered light is unpolarized. For simplicity, also assume that the refracted light is emitted isotropically, although this is not necessary to the argument. Then the
396
14. Polarization
reflection from the surface should be described by the Fresnel reflection coefficients Rx(g/2) and /? (l (g/2), and the volume scattering by the remainder w - Se, so that [wp(g)]±=R±(g/2) + (wFrom equation (14.8) the polarization of the medium should then be given by
. I
P(i
(14.12) where LR{g) = R±{g/2)-R,{g/2)
(14.13)
is the difference between the perpendicular and parallel Fresnel reflection coefficients. Equation (14.12) is plotted in Figure 14.3 as a function of g for an area viewed at e = 60° in the principal plane of a medium of particles whose real refractive index is 1.50 for several values of w. Note that the maximum of P occurs close to 2ftB only when w = Se, that is, Figure 14.3. Theoretical positive-polarization curves of light scattered by a particulate medium of particles of refractive index n = 1.5 if the only polarized component is the light specularly reflected from the particle surfaces. Curves for several different values of the single-scattering albedo are shown. The angle of emergence is e = 60°.
0.2
30 60 g (degrees)
180
397
14.D. The positive branch of polarization
when the particle is completely opaque. As w increases, Px decreases, and gx shifts toward longer wavelengths. The reason for this shift is that the numerator of equation (14.12) is A/?, which does not peak at 2&B, but at much larger phase angles near 160°. A/?(g) is plotted versus g for several values of the refractive index in Figure 14.4. One way to reduce gx is to increase the denominator of (14.12) at large phase angles by making p(g) forward-scattering. However, this would require a highly anisotropic scattering function. Because the particle phase functions of many materials, including the lunar regolith, are backscattering, this explanation is clearly inadequate. The other way in which gx may be reduced is if the forwardtransmitted component of the light is not randomly polarized, but retains some negative polarization, which partially cancels the positively polarized reflection from the surface. This was indeed seen to be the case for the large translucent particles studied in the laboratory (Chapter 6). Additional support for this hypothesis is given in Figure 14.5, which plots the quantity &[wp(g)] for several different size fractions of a powdered olivine basalt. These quantities were obtained as follows: The bidirectional-radiance factors irr(i>, e,g) of the basalt
Figure 14.4. Difference between the two components of the Fresnel reflection coefficient as a function of phase angle for several values of the refractive index. The dashed curve is for the refractive index of metallic iron at A = 0.55 ju.m. Calculated from the values of Yolken and Kruger (1965).
0.8
0.6
AR(g/2) 0.4 0.2
30
60
90 g (degrees)
120
150
180
398
14. Polarization
powders were measured in the principal plane for parallel and perpendicular directions of polarization. Unpolarized light was incident, and e = 60°. These radiance factors are reflectances relative to a Halon standard measured at I = 0 and e = 5°, which was assumed to have a reflectance close to Lambertian, rL = J/v. The surface was assumed to be macroscopically smooth. Then at angles large enough that B(g) is small, equation (14.5) can be written
which may be solved for A[wp(g)], ±
,
e
,
g
)
.
(14.14)
In Figure 14.5a, A[wp(g)] calculated as described earlier is plotted along with the Fresnel difference function A/?(g) for n = 1.7, which is representative of the refractive indices of the minerals in the olivine basalt. If the polarization were due exclusively to Fresnel reflection, then the data for A[vvp(g)] should fall close to the curve of A/?(g). This is indeed the case when g < 80°. However, as the phase angle increases, the points of A[wp(g)] fall below the AZ?(g) curve, evidently because the forward-scattered, once-transmitted light is partly negatively polarized. Shown in Figure 14.5b is a similar analysis for lunar soil. The data for these samples exhibit the same behavior as the fine basalt powder, and the same conclusions follow. According to Dollfus and Bowell (1971), on the lunar surface, gx increases as Px increases. This behavior is consistent with the hypothesis that the transmitted light is partially negatively polarized, because brighter areas are more transparent and hence have smaller Px and gx. Vanderbilt et al. (1985) have published data on the intensity and polarization of light of several wavelengths reflected from wheat. The spectrum of the reflected radiance exhibits the strong absorption bands of chlorophyll at 0.48 and 0.66 jam, but the difference spectrum I± - /|| does not. The authors argue correctly from this observation that only specular reflection contributes to I±—I v However, the largest phase angles in their measurement were about 80°, but the transmitted light is appreciable only at larger phase angles. Alternatively, this vegetation may have too large an optical thickness for the refracted light to have appreciable negative polarization. By contrast,
399
14.D. The positive branch of polarization Figure 14.5. Difference between the two components of polarization versus phase angle measured for several different powders and compared with the difference between the Fresnel reflection coefficients. (Top) Powdered terrestrial olivine basalt of several sizes. (Bottom) Apollo 11 and 12 soil samples. The dashed curves give A Rig) calculated for n = 1.7. 100 0.3
:
50 ^im 30 15 - - AR(n =1.70)
30
60 90 g (degrees)
120
150
—e— Apollo 11
0.3
—'—Apollo 12 1
AR(n = 1.70)
0.2 A[wp(g)] 0.1
30
60
90
g (degrees)
120
150
400
14. Polarization
Woessner and Hapke (1987) have measured the polarization of light scattered by clover and concluded that the light transmitted through this vegetation affects I± — 1^ in a manner similar to that for the silicate materials of Figure 14.5. It may be concluded that for nonopaque particles the amplitude and angle of the maximum of the positive branch of polarization are not determined by the Brewster angle, but by the negatively polarized light that is transmitted through the particles in the forward direction, which becomes important for g > 2&c. The polarization curves have maxima around 100° because the refracted light is an important component of the scattering for larger phase angles. This component is controlled by particle size and by the internal absorption and scattering coefficients. However, for a completely opaque mineral, only surface reflection contributes to the particle scattering, so that P x and gx should be close to the values predicted from the Fresnel equations. The differences between the Fresnel coefficients of materials with large imaginary components of refractive index, such as metals, also peak at very large phase angles (Figure 14.4; also see Figure 14.8). Thus, a value of gx significantly less than 160° is prima facie evidence that the regolith particles are translucent, rather than opaque. The observed equality between A[wp(g)] and Ai?(g) over the range of phase angles between about 30° and 80° suggests that it might be possible to infer the refractive index of the particles of a regolith by measuring the difference between the absolute reflectances in the two directions of polarization at these phase angles. Using the equations of this chapter, this difference can be converted to A[wp(g)] = AR(g/2) which depends only on the refractive index. 14.D.2. The polarization-albedo relation (the Umov effect)
Returning to equation (14.8), for any given values of i, e, and g the denominator increases monotonically with w. In addition, if nonopaque minerals are present, any increase in w due to a decrease in internal absorption will be accompanied by a decrease in the numerator at large phase angles because of the negatively polarized transmitted light. Both effects will cause the amplitude of the positive branch to decrease as w increases. Hence, there is an inverse relation between the amplitude of the positive branch of the polarization curve and the reflectance. This relation is known as the Umov effect, after the
14.D. The positive branch of polarization
401
Russian scientist who first observed and explained it (Umov, 1905). The Umov effect is one of those phenomena that keep getting rediscovered from time to time. The Umov effect is illustrated in Figure 14.1. Before full moon, the eastern hemisphere, which is dominated by brighter highlands, is illuminated and has a lower Px, whereas after full moon, the western hemisphere, whose surface has a large number of dark maria, is illuminated and has a higher Px. Because of the Umov effect, a form of spectroscopy can be carried out by measuring P(g) as a function of A, which has the advantage of requiring the measurement of only a ratio, rather than of an absolute intensity. Two measures of reflectance are the normal albedo An of a resolved surface area and the physical or geometric albedo Ap of the integrated radiance from a body, which is the weighted average of An. Thus, the Umov effect implies that there should be an inverse relation between Px and An or Ap, and this is indeed found to be the case. Unfortunately, although the Umov effect describes a general trend, there is no unique relation that is valid for all materials. There are several reasons for this. First, equation (14.8) shows that Px is determined by p(g) and by the H functions at large phase angles, whereas An and Ap depend on these quantities at g = 0; in general, these functions change in different ways in different materials as w changes. Second, the negative polarization in the transmitted component can be decreased by increasing the amount of internal scattering in the particles, which randomizes the polarization, but may not affect w appreciably. Third, w may also be increased either by decreasing the absorption of the transmitted light, which decreases both the numerator and the denominator of (14.8), or by increasing the real part of the refractive index, which primarily increases the numerator. However, homogeneous classes of materials may possess a quasiunique Umov relation. An example is the lunar regolith, which is relatively homogeneous because it consists of mafic silicate minerals and glasses and is the product of meteorite impacts. Dollfus and Bowell (1971) found that areas on the moon observed telescopically from earth obey the empirical relation log^ll--c1logPx-c2,
(14.15)
where c ^ 0.724 and c 2 —1.81. This is illustrated in Figure 14.6. Other materials, such as certain terrestrial silicate rock powders and
402
14. Polarization
pulverized meteorites, obey similar rules, except that the constants have different numerical values. Geake and Dollfus (1986) have suggested that there is a correlation between c 2 and particle size, although the generality of such a relation is doubtful. 14.D.3. The slope-albedo relation
The inverse relation between the amplitude of the positive branch and the albedo also manifests itself in other properties of P(g) in the positive region. In particular, KenKnight et al. (1967) and Widorn (1967) independently pointed out that the slope ht of the polarization at the inversion angle is inversely related to albedo. This relation is shown in Figure 14.7. Over a fairly large range of albedos, ht and An obey the empirical rule n
=
(14.16)
-c3loghi-c4,
where c 3 ^1.00 and c4-3J7 (Bowell et al., 1973). The physical albedo Ap may be substituted for An in (14.16). However, the curve saturates for An < 0.06, so that there is not a unique relation between albedo and slope for very dark materials. Figure 14.6. Plot of the normal albedo against the maximum polarization for 144 areas on the moon. The line is the empirical function l o g ^ p = -0.724 log Px - 1.81. From Dollfus and Bowell (1971); courtesy of Springer-Verlag. I
I
'
I
1
1
log(An) --0.8
--0.9
- -1.0
- -1.1
--1.2 -0f8
-0.9
-1-0
•
-1.2 I
i
"1.3 I
iog(Px) -1.4 i I
i
-1.5 I
403
14.D. The positive branch of polarization
The slope-albedo relation appears to be unique for a wider range of materials than is the maximum polarization-albedo relation. Probably, the reason is that the numerator of (14.8) is less affected by the transmitted, negatively polarized light at phase angles significantly smaller than gx, and also the quantities in the denominator of (14.8) are closer to their zero phase values at g( than at gx. This rule has important applications in the determination of asteroid albedos and diameters. For many asteroids, gt is small enough that hi can be determined from the earth, whereas the measurement of Px at gx requires a polarimeter on board a spacecraft in the outer solar system. If the albedo of an object can be found from hi9 its size can be calculated from a measurement of absolute integral brightness, Figure 14.7. Log-log plot of normal albedo against the slope of the polarization phase curve at the inversion angle for 95 samples of lunar soil, pulverized terrestrial rocks, and pulverized meteorites. From Geake and Dollfus (1986), courtesy of the Royal Astronomical Society.
.01 — .0002
.0004 .0006
.001
.002 1
h: (degrees" )
.004
404
14. Polarization
even though the object is too small to be resolved (Zellner, Gehrels, and Gradie, 1974; Zellner and Gradie, 1976). 14.E. The negative branch of polarization
The negative branch of the polarization phase curve is one of the enigmas of planetary remote sensing. In spite of the fact that it has been known since the 1920s, it has defied repeated attempts to account for it quantitatively. Yet virtually all pulverized materials display negative polarization at small phase angles. Indeed, the observation by Lyot of a well-developed negative branch of polarization in light reflected from the moon was one of the earliest indications that the lunar surface was covered with a fine-grained regolith. The negative branch cannot be explained in general as a result of single scattering. Although the light forward-refracted through a particle is negatively polarized, transparency is not required. Powders of highly opaque materials, such as metals, display negative polarization (Figure 14.8). In fact, decreasing the transparency of the particles Figure 14.8. Polarization phase curve for two metallic iron powders about 20 fim (crosses) and 3 jLtm (circles) in size. Note that the smaller particles have larger negative polarization. 0.5 0.4
0.3
0.2
0.1
-0.1 20
40
60 80 g (degrees)
100
120
140
14.E. The negative branch ofpolarization
405
decreases / in the denominator of equation (14.2) and increases the polarization, including the amplitude of the negative branch. Moreover, unless the particle is highly regular, such as a sphere, the amount of negative polarization in the forward-scattered light is less than the positive in the surface-reflected light, so that the net polarization is almost always positive. Several observations indicate that the negative branch is caused by multiple scattering: (1) Dollfus (1956; see also Bowell and Zellner, 1974; Dollfus et al, 1989) described experiments in which carbon particles rising in the smoke above a flame did not display negative polarization. However, when those particles were collected into a thick coating on a plate, they had a negative branch. (2) Similarly, when Dollfus (1956) allowed a stream of wellseparated sand grains to fall in front of a polarimeter, there was no negative polarization, but a pile of the same grains had a negative branch. (3) In a particularly revealing experiment, Geake, Geake, and Zellner (1984) placed a single layer of glass particles on a surface of black silicone putty. As shown in Figure 14.9, the layer exhibited negative polarization. The particles were then pushed down into the putty until their tops were nearly flush with the putty. The negative polarization disappeared. The only notable difference between the two media was that pushing the particles into the clay had the effect of separating them with an opaque layer and preventing double scatterings. (4) The amplitude of the negative branch depends strongly on the filling factor in a particulate medium, in the sense that a more closely packed powder has a stronger negative polarization. This is illustrated in Figure 14.10. A commercial silicon carbide abrasive powder with particles about 20 /x,m in size was sedimented in acetone, which was then allowed to evaporate. The sediment had a strong negative branch. Microscopic examination showed that the grains in the sediment were closely packed. However, when the particles were sifted into a low-density deposit, the negative polarization was much weaker. The experiment with the silicon carbide abrasive indicated not only that the efficacy of the mechanism that produces the negative polarization is increased by placing the particles closer together but also
14. Polarization 406 that close packing apparently partially blocks the light transmitted through the particle and increases the positive polarization dramatically. One of the earliest explanations offered for the negative polarization, and the one that underlies most of the models proposed to explain it, is due to Ohman (1955). Ohman pointed out that double specular reflection, in which the intermediate path is perpendicular to the scattering plane, would generate negative polarization. This is illustrated in Figure 14.11. The Fresnel coefficient for ray S-l-A, which is reflected from the surface of particle 1, is larger for the component with its electric vector perpendicular to the scattering plane, and so is positively polarized. However, for the doubly reflected ray S-1-2-B, which reflects at nearly right angles from particles 1 and 2, there is an intermediate scattering plane that is perpendicular to the main scattering plane. Hence, for the doubly scattered ray there Figure 14.9. Polarization phase curves for powdered glass on silicone putty. Open circles, glass grains resting on surface of putty; filled circles, grains pushed into surface; crosses, multiple layers of glass powder. From Geake et al. (1984), courtesy of the Royal Astronomical Society. O 2.0 -
1.6 -
1.2 -
grains on surface
o 0.8
0.4 -
o o
•
O
o o
+ O
-0.4 thick layer of grains -0.8
_ -
grains pushed in
0
-
o
l
i
T
t 8
I
I 12
g (degrees)
I
I 16
14.E. The negative branch of polarization
407
are two Fresnel reflections, each of which produces light that is preferentially polarized parallel to the main scattering plane. That the Ohman mechanism can produce negative polarization is verified by Figure 14.12, which shows the polarization in the light scattered by two copper spheres just touching each other and resting on black velvet. When the line between the centers of the spheres was Figure 14.10. Polarization phase curves of SiC powder about 15 /im in size, showing the effect of porosity. I
•
i
I
O
•
- packed
>
0.5
- sifted
0.4
•
1
•
O .
A
A
A
•
•
.
!
-
-
o 0.1
n
*
o
0.2
0
•
o
0.3
P
•
o
A
A
A
•
A A A A {
;8oo o ooo -
H
20
40
60
80
g (degrees) Figure 14.11.
-Or
100
120
140
408
14. Polarization
parallel to the scattering plane there was only a small amount of negative polarization. Strong negative polarization was observed when the line between the centers was perpendicular to the plane and sideways reflections could occur. A single sphere produced no negative polarization. (However, it is of interest to note that even the parallel configuration produced weak negative polarization.) The difficulty with applying the Ohman mechanism to particulate media concerns the statistical azimuthal symmetry of the positions of the particles. Longitudinal scatterings such as S-1-3-C in Figure 14.11, in which all rays are parallel to the scattering plane, and for which the net polarization is positive, should be just as frequent as transverse scatterings, such as S-1-2-B. Quantitative calculations show that the polarizations caused by the longitudinal and transverse scatterings cancel each other to a high degree, leaving only the small net positive polarization due to single scatterings. Most models to produce negative polarization rely on some hypothesized geometric property of the surface to block the light from one of the longitudinal scatterings. Usually the surface is assumed to be covered with vertical-walled pits or cracks lined with particles (e.g., Wolff, 1975, 1980, 1981; Steigman, 1978; Bandermann, Kemp, and Figure 14.12. Polarization phase curves for a single copper sphere approximately 5 jLim in diameter (crosses), for two copper spheres with the line joining their centers parallel to the scattering plane (open circles), and for two copper spheres with the line perpendicular to the scattering plane (filled circles). - single sphere - two spheres perpendicular - two spheres parallel
ooo -2
10
20
30 g (degrees)
40
50
60
14.E. The negative branch of polarization
409
Wolstencroft, 1972; Shkuratov, 1982; Kolokolova, 1985, 1990). When such a pit is viewed from any direction other than vertical, the side closest to the detector is not visible, so that the light scattered from that surface element does not contribute to the polarization. The Wolff model is the most detailed of these types of models and is a numerical calculation in the form of a FORTRAN computer program. It consists of semiempirical functions containing several parameters that, if properly chosen, can indeed describe the intensity and polarization quite well. The surface is assumed to be covered with pits, and scattering from each wall of a pit is assigned an empirical blocking function. One of the longitudinal scatterings is assigned a larger blocking function than any of the transverse scatterings, thus producing negative polarization. However, the basis for the choice of the values of the various parameters is obscure, and the Wolff model has not been widely accepted. The difficulty with models involving pits is that pits do not seem to be necessary to cause negative polarization. Microscopic examination showed that the surface of the sedimented silicon carbide abrasive shown in Figure 14.10 was quite smooth. Geake et al. (1984) measured the polarization from pits in black silicone putty that had been covered with a layer of glass particles. To be sure, the glass-lined pits produced negative polarization, but so did a simple layer without pits. Some authors have assumed that the interstices between particles would act as pits. However, this is unconvincing. Negative polarization is observed from layers of particles that are equant, convex, and gently rounded in shape. There is a vast difference between the side of such a particle and the sharp vertical wall of a pit. No models that involve pits appear to be adequate to account for the negative polarization from general particulate media. However, negative polarization is observed in light scattered from some samples of volcanic foam and scoria, and the pit models may be appropriate for this type of material. A different type of mechanism was proposed by Hopfield (1966), who pointed out that the light diffracted past a straight edge is partially polarized with the electric vector parallel to the edge. Hopfield considered a particle of square cross section above a diffusely scattering substrate. If viewed from any angle other than zero phase, a shadow is visible under that edge which is nearest the detector and oriented perpendicular to the scattering plane. The light
410
14. Polarization
diffracted from this edge is reduced in intensity, because it comes from the shadow. However, Zellner (Geake et al., 1984) quoted some unpublished experiments he conducted with K. Lumme showing that the negative polarization produced by this mechanism was far too weak to explain the negative branch. A promising mechanism for producing the negative branch of the polarization phase curve was recently proposed independently by Shkuratov (1989) and Muinonen (1990) and involves coherent backscatter. The coherent-backscatter phenomenon was discussed in Section 8.H.4.C. It gives rise to a type of opposition effect from media in which small particles are separated by distances of the order of the wavelength. When the particle separations are this small, parts of the incident wave that are scattered over different paths can interfere coherently with each other. In particular, portions of the incident wave that are multiply reflected between scatterers over the same path, but in opposite directions, combine coherently upon emerging from the medium, and if they are in phase, they produce a peak of increased intensity by constructive interference. Shkuratov (1989) and Muinonen (1990) pointed out that this phenomenon is inherently anisotropic in such a way as to emphasize the transversely scattered portions of the wave. This is illustrated schematically in Figure 14.13. The portions of the incident wave transversely scattered along paths A-1-2-A and B-2-1-B' and emerging in the same direction are in phase for any value of g. However, the portions of the Figure 14.13.
14. E. The negative branch ofpolarization
411
wave longitudinally scattered along paths C-2-3-C and D-3-2-D' are in phase only at g = 0 and will be out of phase if g > A /2TT5?, where Sf is the separation of the scatterers. Hence, the intensity of the negatively polarized, transverse, doubly scattered wave is roughly twice that of the positively polarized, longitudinal, doubly scattered wave, except at phase angles close to zero. Both authors have made quantitative estimates of the size of the effect. Muinonen (1990) calculated the polarization from double scatterings between dipoles of varying separations and position. Shkuratov (1989) assumed that the scatterers reflected light like Fresnel surfaces. Both models produced large amounts of negative polarization. The coherent-backscatter model also accounts for the observation that a medium of small particles has greater negative polarization than one of larger particles. However, several questions must be resolved before this mechanism can be accepted as the general cause of negative polarization in particulate media. One difficulty is that it is not clear that the coherence will also occur in scattering from media composed of particles that are large compared with the wavelength and for which negative polarization is observed. A second problem is that interference phenomena inherently have a strong wavelength dependence. However, Dollfus and Bowell (1971) have measured the polarization of a large number of regions on the moon at several wavelengths between 0.33 and 1.05 jum. Over this interval, in which the wavelength varied by a factor of 3, there was no change in the negative branch. Third, the coherent-backscatter opposition effects observed in the laboratory are less than 1°wide, whereas the negative branches of polarization typically are 23° wide. Fourth, neither model includes the unpolarized light scattered by the particles. Pending further experimental investigation and theoretical development, coherent-backscatter models must be regarded as interesting but unproven speculations. In spite of the lack of theoretical understanding, Dollfus and his co-workers (Geake and Dollfus, 1986; Dollfus et al., 1989) have discovered a number of empirical relations between the size of the negative branch of polarization and the properties of various particulate media. These relations are summarized in Figure 14.14, which is a plot of the amplitude of the polarization minimum Pn versus gi9 the inversion angle. There appear to be three regions, which are labeled 1, 2, and 3 in Figure 14.14. Coarse chunks of terrestrial and lunar
412
14. Polarization
rocks fall in region 1, and finely pulverized terrestrial volcanic rocks and meteorites lie in region 3. Lunar soil occupies region 2. Most asteroids fall in region 3, implying that these bodies are covered with a fine-grained regolith. However, until the negative branch of the polarization phase curve is better understood, the general applicability of these empirical relations remains uncertain. Figure 14.14. Plot of the polarization minimum against the inversion angle, showing the three fields. From Geake and Dollfus (1986), courtesy of the Royal Astronomical Society. 1
1
i
i
i
I.L-1.-1-
1 ' •
i
/
r
N
S
J-\ \
-1.0
N
\
\
\
\ \
\
terrestrial rocks
-
rocks and meteorites
\
1. Lunar and
_
/ /
\
/ *" / __ /
'
\ \ :
: / ]/i
\
I
y
*" -1.5
'1
3. Pulverized
^«.
-0.5
1
-
i i
i; \; \ ;i ^
\
t i i i i i i i i i \ i \ \
2. Lunar
p-
fines 1 __
i \
\
_
\ i\ i
-2.0 — 1
10
1
1
1
|
-r'i
I
15
20 9; (degrees)
i 25
APPENDIXA
A brief review of vector calculus
In this appendix a number of useful relations in vector calculus are given without proof. More detailed discussions can be found in standard texts, such as those by Chorlton (1976) and Lass (1950). Consider two vectors A and B, differing in direction by angle 0:
where u^, u y , and u z are unit vectors in the x, y, and z directions, respectively. Then A and B can be multiplied together in two ways: as a scalar product (or dot product) A-B = \A\ \B\cos 0 = AXBX + AyBy + AZBZ,
(A.I)
or as a vector product (or cross-product) - nx(AyBz
- AzBy) + uy(AzBx
- AxBz) + uz(AxBy
-
AyBx), (A.2)
where n is a unit vector perpendicular to the plane of A and B with direction that a right-handed screw would move if rotated from A to B (Figure A.I). Two useful identities, which can easily be proved by direct substitution, are A X (B X C) « ( A • C)B - (A • B)C
(A.3a)
and A-[BX(CXD)]=(AXB)-(CXD).
413
(A.3b)
414
Appendix A
Both scalar and vector derivatives can be formed by using the differential vector operator "nabla" or "del," which in rectangular coordinates is d and treating it symbolically like a vector quantity, as follows. Let F be a scalar field, F = F(x,y,z), and let G be a vector field, G = uxGx(x,y,z) + uyGy(x>y,z) + uzGz(x,y,z). Then the gradient of the scalar is a vector, defined as dF dF dF ++^. (A.5) For example, the gradient of the electric potential is the electric field Ee. The divergence of a vector is analogous to the scalar multiplication of V and G and is a scalar:
The divergence is a measure of the way fields change as they diverge from a source - for example, electric-field lines around a charge. The curl of a vector is analogous to vector multiplication of V and G and is a vector: curlG =
Curls are associated with rotational or looping behavior - for example, the vortex flow of fluids or magnetic-field lines around a line of Figure A.l. Relation between vectors in a cross-product. A
AXB
A brief review of vector calculus
415
current. The direction of the curl vector is perpendicular to G and parallel to the way the thumb of your right hand would point if you pointed your fingers along G, as shown in Figure A.2. An operator frequently encountered in physical problems is the Laplacian, divgrad, or "del squared." In rectangular coordinates, div • grad = V2 = - ^z + —Tz + T~2dx dy dzz
(A.8)
Representation of these vector operators in other coordinate systems can be found in standard texts. The Laplacian can operate on either a vector or a scalar. Two useful identities, which can be proved by direct substitution, are curl(curlG) = grad(divG) - (divgrad)G, or V x ( V x G ) = V(V-G)-V 2 G,
(A.9)
and div(A X B) = A • curl B - B • curl A, or V (AXB) = A V X B - B V XA.
Figure A.2. Curl of a vector. VXG
(A.10)
416
Appendix A Figure A.3. Gauss's theorem. G
A powerful integral relation, which holds if G is defined over a volume 3^ bounded by a closed surface A, is Gauss's theorem, (A.ll) This states that the integral of the divergence of G over the volume
APPENDIX B
Functions of a complex variable
In this appendix the properties of a function of a complex variable are reviewed briefly. For rigorous proofs and applications of the formulas given here, the reader is referred to standard textbooks (e.g., Paliouras, 1975; Churchill, 1944). Let z be a complex number, z = x + iy, (B.I) where i = y/ — 1. In polar coordinates, z is z = pei0,
(B.2)
where e' 0 = cos0 + Zsin0, (B.3) 2 2 1/2 and p = (JC + y ) , so that Jt = pcos0 and y = psin0. The complex conjugate z* of z is found by replacing i by — i: z* = x - / y . (B.4) Let /(z) be a function of z = JC + /y, with real and imaginary parts u and i;, respectively, f(z) = u(x,y)+iv(x,y), and with complex conjugate f*(z) = u(x,y)-iv(x,y). A function f(z) is analytic at a point if the derivative df/dz exists and is independent of the direction in the complex plane in which the derivative is taken. The conditions for this to be true may be found as follows. Suppose the derivative of /(z) is taken along a direction parallel to the x axis; then df/dz =du/dx+ idv/dx. On the other hand, suppose the derivative is taken parallel to the y axis; then df/dz = (l/i)du/dy + dv/dy. In order that the two derivatives be equal, it 417
418
Appendix B
must be true that du/dx = dv/dy
and dv/dx = —du/dy.
(B.5)
These are the Cauchy-Riemann conditions.
The Cauchy-Goursat theorem states that if /(z) is analytic in some region of the complex plane, then )rfz = 0,
(B.6)
where ^ is any closed curve in the region. A consequence of this theorem is that the integral of /(z),
along any path joining the points zx and z 2 is independent of the specific path as long as it is entirely in the region where /(z) is analytic. Furthermore, the value of / at a point z is determined by the values along any closed curve % surrounding z in the analytic region by Cauchy 's integral formula:
(B.7) If /(z) is analytic throughout the upper half-plane, including the real axis, and also has the property that /(z)->0 as |z|->«>, then (B.7) can be transformed into an integral along the real axis, as follows. Evaluate the integral around the closed curve %* shown in Figure B.I. Z' - plane
Functions of a complex variable
419
Figure B.I. Then
Now, as ^ -> 00 the last integral on the right vanishes, and as A* the second integral on the right becomes - iwfiz). Hence, (B.8) Separating real and imaginary parts of (B.8) gives , , 1
(B.9a)
and 1
(B.9b)
Thus, if /(z) has the properties stated earlier, its real and imaginary parts are not independent, but are related through (B.9). Equations (B.9) can be put in a particularly useful form if, in addition to being analytic in the upper half-plane and vanishing at infinity, /(z) possesses the property of crossing symmetry; that is, / ( - z ) = /*(z), so that w ( - z ) = w(z) and v(-z)=-v(z). Then (B.9a) becomes
T4 and, similarly, (B.9b) is
Z'2-Z
APPENDIX C
The wave equation in spherical coordinates
In this appendix the general solution to the wave equation in spherical coordinates using the separation-of-variables method is reviewed. For further discussions, the many excellent textbooks on mathematical methods in physics should be consulted, such as Arfken (1985), Morse and Feshbach (1953), or Margenau and Murphy (1956). The wave equation (Chapter 2) is V 2 F - ( 1 / v2)d2F/dt2
= 0,
(C.I)
where F is the field quantity, v is the propagation velocity, and t is the time. The Laplacian in spherical coordinates is
-2F=1_JL(
2^\ , _J_A_(
. qdF\
*d,V 9] V i # ^ l S m
1 8 2F
?#] V i 2 # # 2 ' (C.2)
where A* is the radial coordinate, # is the polar angle, and if/ is the azimuth angle (Figure C.I). In a very large class of problems it is possible to write the solution to the wave equation as the product of separate functions of the independent variables: F(r9#,4,,t)-R(r)e(#)QMT(t).
(C.3)
(The reader should be cautioned that this does not always work. Not infrequently one will obtain an apparent solution to a differential equation, only to find that it is not possible to satisfy the boundary conditions.) It is almost always found that more than one solution of the form (C.3) exists (usually an infinite number of them), so that the 420
The wave equation in spherical coordinates
All
complete solution to (C.I) consists of a sum of terms of the form of (C.3). To solve (C.I) using the separation-of-variables method, insert (C.3) into (C.I), divide by F, and multiply by v2 to obtain
J
1
d •
dif,
(
-
2
1 d2T T dt2'
(C.4)
Now, the left-hand side of (C.4) is solely a function of the spatial variables A% #, and $, while the right-hand side is a function of t only. Because these variables are independent, the only way the equality can hold is if both sides are equal to some constant. Call this separation constant 3tx. Equating the right-hand side of (C.4) to Jft gives
rf2r/^2-^r
= 0.
(C.5)
Equating the left-hand side of (C.4) to Xx and rearranging gives
*2# d I 7dR\ . sind d ( . «d©\
1 d 2® * dty2 ' (C.6)
Figure C.I. Spherical coordinates.
422
Appendix C
Because each side of (C.6) is a function of different independent variables, both sides must equal a constant. Call this second separation constant -3t2. Equating the right side to X2 gives
±p; + JT2O = 0.
(C.7)
Setting the left side of (C.6) equal to JF2 gives Xx
2
1 d I 2dR\
&2
1
d
^ '
(C8)
The same argument may again be made that both sides of (C.8) must equal a constant, say Xv Equating the right side of (C.8) to this constant gives (C9)
and for the left side,
This completes the separation of the wave equation into four separate differential equations for functions of the independent variables. The values of the separation constants must be determined by various boundary and symmetry conditions of the specific problem. The next step is the solution of the separated differential equations. The solutions of (C.5) are of the form exp(±^3fj"f), as may be verified by direct substitution. As discussed in Chapter 2, in this book we are concerned with waves that vary periodically with time as e~2iriv\ where v is the frequency. These two expressions will be of the same form if jr1=-47rV.
(C.ll)
Then the solution of interest for T is where To is a constant that must be determined by the boundary conditions. Note that we could have obtained the same result simply by assuming that the wave equation has periodic solutions of the form F(A*,^,i/r,O = ^(^,^,(/r)e" 2 7 r ^ and substituting this into (C.I) to
The wave equation in spherical coordinates
423
obtain
where and A is the wavelength. If 5?= 0, equation (C.13) is called Laplace's equation, and if
(C.16)
and the AJs and Bm's are constants that must be determined by the boundary conditions. The general solutions to equation (C.9) for & are the set of functions known as confluent hypergeometric functions. They have the property that they are infinite at # = 0 and ir unless 3F3 has the special value
jr3=/(^+i),
(en)
where / is an integer > rn. Because most problems of interest include 0 and IT, the solutions with singularities are not physically acceptable, and the separation constant must be chosen to have the value given by (C.17). The solutions of (C.9) with X2 = m2 and JT3= / ( / +1) are the associated Legendre functions, ), (C.18)
424
Appendix C
where Eim is a constant. These functions are tabulated in many places (e.g., Jahnke and Emde, 1945; Abramowitz and Stegun, 1965). If the second index is omitted, the value of m is understood to be 0 (i.e., Pe = P / o ), and the functions are called simply Legendre polynomials. The four lowest Legendre polynomials are plotted in Figure C.2. The first few associated Legendre functions are as follows: P 20 = /> 2 = ± ( 3 c o s 2 0 - 1) ^30 = ^3 = | ( 5 c o s 3 0 - 3 c o s 0 ) Pn = sin 0
P2l = 3 cos 0 sin 0
P3l = |(sin 0 + 5 sin 30)
P22 = 3 sin20
P32 = 15 cos 0 sin20 P 33 = 15sin30
If JTt == 0, the solution to (CIO) is of the form i ? (^) = Q / - h D ^ - ( / + 1 ) ,
(C.19a)
where Ce and Dg are constants that must be determined from the boundary conditions. However, if 3tx = 0, then v = 0, so that this solution for i?(/0 applies only to static situations. Because such
Figure C.2. First four Legendre polynomials. , T S J '
0.5
• • • • • • • • • •
-
/
\ r
i
o
\ :
-0.5
\
v
\
x
:
:
V ^
v
\
/
\
/
•
\
X
•
-1 30
60
90 120 g (degrees)
150
180
The wave equation in spherical coordinates
425
problems cannot involve propagating waves, they will not be considered further here. The solutions to (CIO) with - Xx / v2 = JT 2 # 0 and Jf3 = / ( I +1) are
*( ') - [Q W
^ + DfY^JTrJlfr,
(C.19b)
where Jp{z) is a Bessel function of the first kind of argument z and order p, and Yp(z) is a Bessel function of the second kind of argument z and order p. Bessel functions of the second kind are also called Neumann functions. As usual, the constants Cf and Df must be determined by the boundary conditions. The functions that result when Bessel functions of half-integral order are divided by {z are called spherical Bessel functions because they occur in the solutions of Helmholtz's equation in spherical coordinates. Thus, the spherical Bessel functions are defined as
*y>
(c-20a)
)-
(C20b)
The Bessel functions are tabulated and discussed in many places (e.g., Watson, 1958; Jahnke and Emde, 1945; Abramowitz and Stegun, 1965). Jahnke and Emde are particularly helpful because the Legendre functions and Bessel functions are displayed graphically in several ways, which allows the student to rapidly grasp their properties. The Bessel functions are oscillatory functions whose amplitude generally decreases with increasing argument. In particular, / 0 (z) = (sin z ) / z and yo(z) = -(cosz)/z. Just as solutions to various problems often are expedited by combining trigonometric functions into complex functions, such as (cos # + /sin#), it is frequently convenient to do the same with the Bessel functions. Thus, Bessel functions of the third kind, or Hankel functions, are defined as = Jp(z) + iYp(zy,
(C.21a.)
= Jp(z)-iYp(z).
(C.21b)
426
Appendix C
Similarly, the spherical Hankel functions are defined as
W:
(c-22a)
)(z).
(C.22b)
For any given value of ( there will be I +1 possible values of m (m goes from 0 to (). Because / can be any integer > 0, the general solution to the wave equation that is periodic in time will be a sum of the various solutions for R( A<)0(#)4>(i/f) for all possible values of t and m. Hence (absorbing To and E(m into the other constants), the general solution to the wave equation is oo
t
F(,>,#,<M)= E E [C,j,(3Tr) +
Dtyti^P^costi)
e=om=o X [Afmcos mil/ + fi^sin m4]e~2tirivt. (C.23) If v = 0, then the Bessel functions are replaced by / and /*~(/f+1), equation (C.19a). Because the associated Legendre functions and the trigonometric functions always appear multiplied together in the solutions to the wave equation in spherical coordinates, their products are called spherical harmonics: ¥ # ( 0 , ^ ) = P/Jcofi #)cos m$
(C.24a)
and ¥ # ( 0 , iff) = />,Jcos #)sin m# (C.24b) are the even and odd spherical harmonic functions, respectively. Thus, the general solution to the wave equation may be written in a form equivalent to (C.23):
F(r,0,0,O= E E [CthfPW + DthfiXr)] e=o^=o x[^/^#(0,*) + ^¥#(0,*)]e- 2 '^, (C.25) where the various constants must be found from the boundary conditions. The physical meaning of writing the solution in the form of (C.25) may be seen from the asymptotic properties of the spherical Hankel 2 functions. When
The wave equation in spherical coordinates
All
and
The first expression represents a wave propagating radially outward whose amplitude falls off as A* ~\ that is, whose power falls off as r ~2. Similarly, the second represents a wave propagating radially inward. The part of (C.24) proportional to the spherical harmonics describes how the strengths of these spherical waves vary with direction. The spherical Bessel functions and spherical harmonic functions are members of a large class of functions that are known as complete orthogonal sets. Let fp(z) be a member of a set of functions of argument z and order p, defined over the interval a
where the Ap's are constants. The set is orthogonal if I constant =£ 0 if p = p'. A typical example is provided by the familiar set of trigonometric functions. The ability of a sum of sine and cosine functions to represent any periodic function is well known and forms the basis for the technique of Fourier analysis and synthesis. That these functions are orthogonal can be illustrated by the following. Let fp(z) = cos pz, a = 0, and b = 2TT, and let p and p' be integers. Then 2TT
0
cos pz cos p'zdz
=f Jf\
Tj[cos(p + pf)z +cos(p-//)z] dz = I ^
77 II P — P
, f
APPENDIX D
Table of symbols
a aE a Aa A Ab Ae Ah An Ap As AL AM sf b B & BQ Bm BRDF c c0 C g7 AC d d D
mean particle radius, a = particle extinction radius, aE surface, area sensitive area of detector area bolometric albedo effective absorbance hemispherical or directional-hemispherical albedo normal albedo physical or geometric albedo spherical or bihemispherical albedo Lambert albedo Minnaert albedo arbitrary constant coefficient in particle phase function opposition effect arbitrary constant amplitude of opposition effect magnetic field bidirectional-reflectance distribution function coefficient in particle phase function speed of light specific heat arbitrary constant relative band contrast differential operator partial differential operator mean particle diameter, D = 2a 428
Table of symbols 3f De e e E E % Ee %e / f / fj F SF g G 3?v &2 h & h0 H Hj Hm %* i I / ID It I2 / ie j / k k0 & K X Ke Ker Kei I
mean ray path length through particle arbitrary constant, distance electric displacement charge of electron, also base of natural logarithms zenith angle of emergence or viewing volume extinction coefficient particle emissivity factor volume emittance factor electric field spatially dependent part of electric field arbitrary function photometric function remission function oscillator strength of /th transition volume emission coefficient source function phase angle volume angular scattering coefficient factor in R with complex m angular-width parameter of opposition effect height characterizing roughness Planck's constant H function moment of H function magnetic intensity scale height yj — 1 zenith angle of incidence radiance radiance at detector viewing surface of scattering medium upwelling radiance downwelling radiance subscript denoting different types electric-current density arbitrary constant irradiance imaginary part of refractive index Boltzmann's constant arbitrary constant volume absorption coefficient roughness correction factor to integral phase function dielectric constant, Ke = Ker + Kei real part of Ke imaginary part of Ke distance, also arbitrary constant
429
430 L 3* m m J? -*k
Mm
71
n N Ne NE No jr
p p Pe
Pe P & PA PD PE
Ps Q
QA QE
Qs
Qs r
r
r r
B
r r
h
hd
rL *& rM r
R
R
Appendix D luminance latitude distance, thickness refractive index, m = n + ik arbitrary constant number bulk density of /th type of material magnetization real part of refractive index arbitrary constant number of particles per unit volume number of free electrons per unit volume effective number of particles per unit volume Avogadro's number total number of particles volume angular-scattering function particle angular-scattering function electric-dipole moment power emitted by dipole polarization factor in derivation of r absorbed power power at detector dipole moment per unit volume extinguished power scattered power emitted power phase integral absorption efficiency extinction efficiency scattering efficiency excluding diffraction scattering efficiency including diffraction bidirectional reflectance distance vector position diffusive reflectance of bottom layer directional-hemispherical reflectance hemispherical-directional reflectance internal diffusive reflectance inside particle Lambert reflectance diffusive reflectance of lower layer Minnaert reflectance diffusive reflectance bidirectional reflectance of a rough-surface medium diffusive reflectance of upper layer specular Fresnel reflection coefficient
Table of symbols
431
& R± /?H RADF REFF s 4 5 Se Si So S t te tt t0 T Tb !T up, ux,
radius of planet perpendicular specular reflection coefficient parallel specular reflection coefficient radiance factor reflectance factor distance, position, line element near-surface internal scattering coefficient volume scattering coefficient total Fresnel reflectivity for externally incident light total Fresnel reflectivity for internally incident light solar constant shadowing function time transmission probability for emergent radiance transmission probability for incident radiance transmittance of layer temperature brightness temperature Fresnel transmission coefficient unit vectors
U % v vV jr w w W W W x AJC X x y f Y
Planck black-body radiation function roughness correction factor to the physical albedo velocity particle volume Stefan-Boltzmann integral black-body radiation function volume volume single-scattering albedo particle single-scattering albedo volume espat function, W = (1 - w)/w particle espat function molecular weight variable, often denoting horizontal position short path length through particle size parameter arbitrary constant variable, often denoting horizontal position y/a/ia+6) emergent radiant power per unit solid angle per unit area of the surface of the medium thermal-emission-function scale height variable, often denoting vertical position average distance between particle, Z = N~l/3 factor in h absorption coefficient
y z Z 2? a
432
Appendix D
ae electric polarizability )8 hemispherical asymmetry factor X average cosine of the surface tilt angle electric susceptibility Xe magnetic susceptibility Xm 8 increment, phase S(x) Dirac delta function of argument x A increment tan A loss tangent Laplacian operator, V 2 = div • grad V2 e particle emissivity s wavelength-averaged emissivity directional emissivity ed se electric permittivity permittivity of free space ee0 hemispherical emissivity eh <j> filling factor integral phase function albedo factor, y = Vl — w y T relative reflectance amplitude of complex specular reflection coefficient at normal 7] incidence if/ phase, also azimuth ^iie spherical harmonic A wavelength A luminance longitude fJL
fJL =
\im fim0 /x 0 v vp
magnetic permeability permeability of free space /x 0 = cos I frequency, also exponent of particle size distribution plasma frequency pi Poynting vector scattering angle particle internal-transmission coefficient mass density electric-charge density particle cross-sectional area particle absorption cross section electrical conductivity particle extinction cross section Stefan-Boltzmann constant particle scattering cross section summation operator
IT
II 6 © p pe a aA
COS e
Table of symbols r r0 v T q> A
optical depth optical thickness Minnaert index angle of refracted ray in scattering function of sphere directionally averaged radiance difference between upwelling and downwelling radiances phase, photometric function acceptance solid angle of detector direction increment of solid angle cosine asymmetry factor electron collision frequency angle Brewster angle critical angle C = \/l — j8w, also azimuth angle
433
BIBLIOGRAPHY
Abeles, F. (1966). Optical Properties and Electronic Structure of Metals and
Alloys, Amsterdam: North Holland. Abramowitz, M., and Stegun, I. (1965). Handbook of Mathematical Functions. New York: Dover Press. Adams, C, and Kattawar, G. (1978). Radiative transfer in spherical shell atmospheres. I. Rayleigh scattering. Icarus, 35, 139-51. Adams, J. (1975). Interpretation of visible and near-infrared diffuse reflectance spectra of pyroxenes and other rock-forming minerals. In Infrared and Raman Spectroscopy of Lunar and Terrestrial Minerals, ed.
C. Karr (pp. 91-116). New York: Academic Press. Adams, J., and Felice, A. (1967). Spectral reflectance 0.4 to 2.0 microns of silicate rock powder. /. Geophys. Res., 72, 5705-15. Allen, C. (1946). The spectrum of the corona at the eclipse of 1940 October 1. Proc. Roy. Astron. Soc. London, 106, 137-50. Ambartsumian, V. (1958). The theory of radiative transfer in planetary atmospheres. In Theoretical Astrophysics, ed. V. Ambartsumian (pp. 550-64). New York: Pergamon Press. Arfken, G. (1985). Mathematical Methods for Physicists. New York: Academic Press. Arnold, G., and Wagner, C. (1988). Grain size influence on the mid-infrared spectra of the minerals. Earth, Moon and Plan., 41, 163-72. Aronson, J., and Emslie, A. (1973). Spectral reflectance and emittance of particulate materials. 2. Application and results. Appl. Opt., 12, 2573-84. (1975). Applications of infrared spectroscopy and radiative transfer to earth sciences. In Infrared and Raman Spectroscopy of Lunar and Terrestrial Minerals, ed. C. Karr (pp. 143-64). New York: Academic Press. Aronson, J., Emslie, A., Ruccia, F., Smallman, C, Smith, E., and Stong, P. (1979). Infrared emittance of fibrous materials. Appl Opt., 18, 2622-33. 434
Bibliography
435
Asano, S., and Yamamoto, G. (1975). Light scattering by a spheroidal particle. Appl. Opt., 14, 29-49. Bandermann, L., Kemp^ J., and Wolstencroft, R. (1972). Circular polarization of light scattered from rough surfaces. Mon. Not. Roy. Astron. Soc, 158, 291-304. Barber, P., and Yeh, C. (1975). Scattering of electromagnetic waves by arbitrarily shaped dielectric bodies. Appl. Opt., 14, 2864-77. Beckmann, P. (1965). Shadowing of random rough surfaces. IEEE Trans. Antennas and Propagation, 13, 384-8. Berreman, D. (1970). Resonant reflectance anomalies: effect of shapes of surface irregularities. Phys. Rev., Bl, 381-9. Bevington, P. (1969). Data Reduction and Error Analysis for the Physical Sciences. New York: McGraw-Hill. Blevin, W., and Brown, W. (1967). Effect of particle separation on the reflectance of semi-infinite diffusers. /. Opt. Soc. Amer., 57, 129-34. Bloss, F. (1961). An Introduction to the Methods of Optical Crystallography. Philadelphia: Holt, Rinehart & Winston. Bobrov, M. (1962). Generalization of the theory of the shadow effect on Saturn's rings to the case of particles of unequal size. Sou. Astron. Astrophys. J., 5, 508-16. Bohren, C. (1986). Applicability of effective-medium theories to problems of scattering and absorption by nonhomogeneous atmospheric particles. /. Atmos. Sci., 43, 468-75. Bohren, C, and Huffman, D. (1983). Absorption and Scattering of Light by Small Particles. New York: Wiley. Borel, C, Gerstl, S., and Powers, B. (1991). The radiosity method in optical remote sensing of structured 3-D surfaces. Rem. Sens. Environ., 36, 13-44. Born, M., and Wolf, E. (1980). Principles of Optics, 6th ed. New York: Pergamon Press. Bottcher, C. (1952). Theory of Electric Polarization. Amsterdam: Elsevier. Bowell, E., Dollfus, A., Zellner, B., and Geake, J. (1973). Polarimetric properties of the lunar surface and its interpretation. Part 6: Albedo determinations from polarimetric measurements. In Proc. 4th Lunar Sci. Conf, ed. W. Gose (pp. 3167-74). New York: Pergamon Press. Bowell, E., Hapke, B., Domingue, D., Lumme, K., Peltoniemi, J., and Harris, A. (1989). Applications of photometric models to asteroids. In Asteroids II, ed. R. Binzel, T. Gehrels, and M. Matthews (pp. 524-56). Tucson: University of Arizona Press. Browell, E., and Anderson, R. (1975). Ultraviolet optical constants of water and ammonia ices. /. Opt. Soc. Amer., 65, 919-26. Bowell, E., and Lumme, K. (1979). Polarimetry and magnitudes of asteroids. In Asteroids, ed. T. Gehrels (pp. 132-69). Tucson: University of Arizona Press.
436
Bibliography
Bowell, E., and Zellner, B. (1974). Polarizations of asteroids and satellites. In Planets, Stars and Nebulae Studied with Photopolarimetry, ed. T. Gehrels (pp. 381-404). Tucson: University of Arizona Press. Brown, R., and Cruikshank, D. (1983). The Uranian satellites: surface compositions and opposition brightness surges. Icarus, 55, 83-92. Brown, R., and Matson, D. (1987). Thermal effects of insolation propagation in the regoliths of airless bodies. Icarus, 72, 84-94. Bruggeman, D. (1935). Berechnung verschiedener physikalischer Konstanten von heterogen Substanzen. I. Dielectrizitatskonstanten und Leifahigkeiten der Mischkorper aus isotropen Substanzen. Ann. Phys. (Leipzig), 24, 636-79. Buratti, B. (1985). Application of a radiative transfer model to bright icy satellites. Icarus, 61, 208-17. Buratti, B., and Veverka, J. (1985). Photometry of rough planetary surfaces: the role of multiple scattering. Icarus, 64, 320-8. Burns, R. (1970). Mineralogical Applications of Crystal Field Theory. Cambridge University Press. Burns, R., Nolet, D., Parkin, K., McCammon, C, and Schwartz, K. (1980). Mixed-valence minerals of iron and titanium: correlations of structural, Mossbauer and electronic spectral data. In Mixed Valence Compounds, ed. D. Brown (pp. 295-336). Boston: Reidel. Camillo, P. (1987). A canopy reflectance model based on an analytical solution to the multiple scattering equation. Rem. Sens. Environ., 23, 453-77. Campbell, M, and Ulrichs, J. (1969). Electrical properties of rocks and their significance for lunar radar absorptions. /. Geophys. Res., 74, 5867-81. Chamberlain, J., and Smith, G. (1970). Interpretation of the Venus CO 2 absorption bands. Astrophys. J., 160, 755-65. Chandrasekhar, S. (1960). Radiative Transfer. New York: Dover. Chapman, S. (1931). The absorption and dissociative or ionizing effect of monochromatic radiation in an atmosphere on a rotating Earth. II. Grazing incidence. Proc. Phys. Soc, 43, 483-501. Chiapetta, P. (1980). A new model for scattering by irregular absorbing particles. Astron. Astrophys., 83, 348-53. Chorlton, F. (1976). Vector and Tensor Methods. New York: Wiley. Churchill, R. (1944). Modern Operational Mathematics in Engineering. New York: McGraw-Hill. Chylek, P., Grams, G., and Pinnick, R. (1976). Light scattering by irregular randomly oriented particles. Science, 193, 480-2. Clark, R. (1983). Spectral properties of mixtures of montmorillonite and dark carbon grains: implications for remote sensing minerals containing chemically and physically adsorbed water. /. Geophys. Res., 88, 10635-44. Clark, R., Kierein, K., and Swayze, G. (1993). Experimental verification of the Hapke reflectance theory. 1: Computation of reflectance as a
Bibliography
437
function of grain size and wavelength based on optical constants. Preprint. Clark, R., and Lucey, P. (1984). Spectral properties of ice-particulate mixtures and implications for remote sensing. 1. Intimate mixtures. /. Geophys. Res., 89, 6341-8. Clark, R., and Roush, T. (1984). Reflectance spectroscopy: quantitative analysis techniques for remote sensing applications. /. Geophys. Res., 89, 6329-40. Cohen, A., and Janezic, G. (1983). Relationships among trapped hole and trapped electron centers in oxidized soda-silica glasses of high purity. Phys. Stat. Sol (a), 77, 619-24. Conel, J. (1969). Infrared emissivities of silicates: experimental results and a cloudy atmosphere model of spectral emission from condensed particulate mediums. /. Geophys. Res., 74, 1614-34. Coulson, K. (1971). The polarization of light in the environment. In Planets, Stars and Nebulae Studied with Photopolarimetry, ed. T. Gehrels (pp. 444-71). Tucson: University of Arizona Press. Cox, C, and Munk, W. (1954). Measurement of the roughness of the sea surface from photographs of the sun's glitter. /. Opt. Soc. Amer., 44, 838-50. Danjon, A. (1949). Photometrie et colorimetrie des planets Mercure et Venus. Bull. Astron., 14, 315-17. Dexter, D. (1956). Absorption of light by atoms in solids. Phys. Rev., 101, 48-55. Dickinson, R., Pinty, B., and Verstraete, M. (1990). Relating surface albedos in GCM to remotely sensed data. Agricul, Forest Meteor., 52, 109-31. Dollfus, A. (1956). Polarisation de la lumiere renvoyee par les corps solides et les nuages naturels. Ann. Astrophys., 19, 83-113. (1961). Polarization studies of planets. In Planets and Satellites, ed. G. Kuiper and B. Middlehurst (pp. 343-99). Chicago: University of Chicago Press. (1962). The polarization of moonlight. In Physics and Astronomy of the Moon, ed. Z. Kopal (pp. 131-60). New York: Academic Press. Dollfus, A., and Bowell, E. (1971). Polarimetric properties of the lunar surfafce and its interpretation, part I. Telescopic observations. Astron. Astrophys., 10, 29-53. Dollfus, A., Wolff, M., Geake, J., Lupishko, D., and Dougherty, L. (1989). Photopolarimetry of asteroids. In Asteroids II, ed. R. Binzel, T. Gehrels, and M. Matthews (pp. 594-615). Tucson: University of Arizona Press. Domingue, D., and Hapke, B. (1989). Fitting theoretical photometric functions to asteroid phase curves. Icarus, 74, 330-6. Domingue, D., Hapke, B., Lockwood, G., and Thompson, D. (1991). Europa's phase curve: implications for surface structure. Icarus, 90, 30-42. Drude, P. (1959). Theory of Optics. New York: Dover Press.
438
Bibliography
Dwight, H. (1947). Tables of Integrals and Other Mathematical Data. New York: Macmillan. Egan, W. (1985). Photometry and Polarization in Remote Sensing. New York: Elsevier. Egan, W., and Hilgeman, T. (1976). Retroreflectance measurements of photometric standards and coatings. Appl Opt., 15, 1845-9. (1978). Spectral reflectance of particulate materials: a Monte Carlo model including asperity scattering. Appl Opt., 17, 245-52. (1979). Optical Properties of Inhomogeneous Materials. New York: Academic Press. Elliot, R. (1966). Electromagnetics. New York: Academic Press. Emslie, A., and Aronson, J. (1973). Spectral reflectance and emittance of particulate materials. 1. Theory. Appl Opt., 12, 2563-72. Esposito, L. (1979). Extensions to the classical calculation of the effect of mutual shadowing in diffuse reflection. Icarus, 39, 69-80. Evans, J. (1962). Radio echo studies of the moon. In Physics and Astronomy of the Moon, ed. Z. Kopal (pp. 429-80). New York: Academic Press. Evans, J., and Hagfors, T. (1968). Radar Astronomy. New York: McGraw-Hill. (1971). Radar studies of the moon. In Advances in Astronomy and Astrophysics, Vol. 8, ed. Z. Kopal (pp. 29-107). New York: Academic Press. Fairchild, M., and Daoust, D. (1988). Goniospectrophotometric analysis of pressed PTFE powder for use as a primary transfer standard. Appl Opt, 27, 3392-6. Fowler, W. (1968). Physics of Color Centers. New York: Academic Press. Fredricksson, K., and Keil, K. (1963). The light-dark structures in the Pantar and Kapoeta stone meteorites. Geochim., Cosmochim. Ada, 27, 717-39. Frohlich, H. (1958). Theory of Dielectrics, 2nd ed. London: Oxford University Press. Fung, A., and Ulaby, F. (1983). Matter-energy interactions in the microwave region. In Manual of Remote Sensing, ed. D. Simonett (pp. 115-64). Falls Church, VA: American Society of Photogrammetry. Gaffey, M., Bell, J., and Cruikshank, D. (1989). Reflectance spectroscopy and asteroid surface mineralogy. In Asteroids II, ed. R. Binzel, T. Gehrels, and M. Matthews (pp. 98-127). Tucson: University of Arizona Press. Galileo (1638). Dialogue on the Great World Systems, trans. G. De Santillana (1953). Chicago: University of Chicago Press. Garbuny, M. (1965). Optical Physics. New York: Academic Press. Geake, J., and Dollfus, A. (1986). Planetary surface texture and albedo from parameter plots of optical polarization data. Mon. Not. Roy. Astron. Soc, 218, 75-91. Geake, J., Geake, M., and Zellner, B. (1984). Experiments to test theoretical models of the polarization of light by rough surfaces. Mon. Not. Roy. Astron. Soc, 210, 89-112.
Bibliography
439
Gehrels, T. (1974). Introduction and overview. In Planets, Stars and Nebulae Studied with Photopolarimetry, ed. T. Gehrels (pp. 3-44). Tucson: University of Arizona Press. Gehrels, T., Coffeen, D., and Owings, D. (1964). Wavelength dependence of polarization. III. The lunar surface. Astron. J., 69, 826-52. Gehrels, T., and Teska, T. (1963). The wavelength dependence of polarization. Appl Opt, 2, 67-77. Gerstl, S., and Zardecki, A. (1985a). Discrete-ordinates finite-element method for atmospheric radiative transfer and remote sensing. Appl. Opt, 24, 81-93. (1985b). Coupled atmosphere/canopy model for remote sensing of plant reflectance features. Appl. Opt, 24, 94-103. Goguen, J. (1981). A Theoretical and Experimental Investigation of the Photometric Functions of Paniculate Surfaces. Ph.D. thesis, Cornell University. Goody, R. (1964). Atmospheric Radiation. I. Theoretical Basis. Oxford University Press. Gradie, J., and Veverka, J. (1982). When are spectral reflectance curves comparable? Icarus, 49, 109-19. Greenberg, J. (1974). Some examples of exact and approximate solutions in small particle scattering: a progress report. In Planets, Stars and Nebulae Studied with Photopolarimetry, ed. T. Gehrels (pp. 107-34). Tucson: University of Arizona Press. Grum, F., and Luckey, G. (1968). Optical sphere paint and a working standard of reflectance. Appl. Opt, 7, 2289-94. Hagfors, T. (1964). Backscatter from an undulating surface with applications to radar returns from the moon. /. Geophys. Res., 69, 3779-84. (1968). Relations between rough surfaces and their scattering properties as applied to radar astronomy. In Radar Astronomy, ed. J. Evans and T. Gehrels (pp. 187-218). New York: McGraw-Hill. Hameen-Anttila, K. (1967). Surface photometry of the planet Mercury. Ann. Acad. Sci. Fenn., Ser., A6, 252, 1-19. Hansen, J., and Arking, A. (1971). Clouds of Venus: evidence for their nature. Science, 171, 669-72. Hansen, J., and Travis, L. (1974). Light scattering in planetary atmospheres. Space Sci. Rev., 16, 527-610. Hapke, B. (1963). A theoretical photometric function for the lunar surface. /. Geophys. Res., 68, 4571-86. (1968). On the particle size distribution of lunar soil. Planet Space Sci., 16, 101-10. (1971). Optical properties of the lunar surface. In Physics and Astronomy of the Moon, ed. Z. Kopal (pp. 155-211). New York: Academic Press. (1981). Bidirectional reflectance spectroscopy. 1. Theory. /. Geophys. Res., 86, 3039-54.
440
Bibliography
(1984). Bidirectional reflectance spectroscopy. 3. Correction for macroscopic roughness. Icarus, 59, 41-59. (1986). Bidirectional reflectance spectroscopy. 4. Extinction and the opposition effect. Icarus, 67, 264-80. (1990). Coherent backscatter and the radar characteristics of outer planet satellites. Icarus, 88, 407-17. Hapke, B., and Blewett, D. (1991). Coherent backscatter model for the unusual radar reflectivity of icy satellites. Nature, 352, 46-7. Hapke, B., Cassidy, W., and Wells, E. (1975). Effects of vapor phase deposition processes on the optical, chemical and magnetic properties of the lunar regolith. The Moon, 13, 339-53. Hapke, B., and Nelson, R. (1975). Evidence for an elemental sulfur component of the clouds from Venus spectrophotometry. /. Atmos. Res., 32, 1211-18. Hapke, B., and Van Horn, H. (1963). Photometric studies of complex surfaces with applications to the moon. /. Geophys. Res., 68, 4545-70. Hapke, B., and Wells, E. (1981). Bidirectional reflectance spectroscopy. 2. Experiments and observations. /. Geophys. Res., 86, 3055-60. Hapke, B., Wells, E., and Wagner, J. (1981). Far-UV, visible and near-IR reflectance spectra of frosts of H 2 O, CO2, NH 3 and SO2. Icarus, 47, 361-7. Hapke, B., and Williams, A. (1988). Search for anomalous opposition spike in crystalline powders. Bull. Amer. Astron. Soc, 20, 808. Harris, D. (1957). Diffuse reflection from planetary atmospheres. Astrophys. J., 126, 408-12. Helfenstein, P. (1986). Derivation and Analysis of Geological Constraints on the Emplacement and Evolution of Terrains on Ganymede from Applied Differential Photometry. Ph.D. thesis, Brown University. (1988). The geological interpretation of photometric surface roughness. Icarus, 73, 462-81. Helfenstein, P., and Veverka, J. (1987). Photometric properties of lunar terrains derived from Hapke's equation. Icarus, 72, 343-57. (1989). Physical characterization of asteroid surfaces from photometric analysis. In Asteroids II, ed. R. Binzel, T. Gehrels, and M. Matthews (pp. 557-93). Tucson: University of Arizona Press. Helfenstein, P., Veverka, J., and Thomas, P. (1988). Uranus satellites: Hapke parameters from Voyager disk-integrated photometry. Icarus, 78, 231-9. Henyey, C, and Greenstein, J. (1941). Diffuse radiation in the galaxy. Astrophys. J., 93, 70-83. Herbst, T., Skrutskie, M., and Nicholson, P. (1987). The phase curve of the Uranian rings. Icarus, 71, 103-14. Hisdal, B. (1965). Reflectance of perfect diffuse and specular samples in the integrating sphere. /. Opt. Soc. Amer., 55, 1122-8.
Bibliography
441
Hodkinson, J. (1963). Light scattering and extinction by irregular particles larger than the wavelength. In Electromagnetic Scattering, ed. M. Kerker (pp. 87-100). New York: Macmillan. Hodkinson, J., and Greenleaves, I. (1963). Computations of light scattering and extinction by spheres according to diffraction and geometrical optics, and some comparisons with the Mie theory. /. Opt. Soc. Amer., 53, 577-88. Holland, A., and Gagne, G. (1970). The scattering of polarized light by polydisperse systems of irregular particles. Appl Opt, 9, 1113-21. Hopfield, J. (1966). Mechanism of lunar polarization. Science, 151, 1380-1. Huguenin, R., and Jones, J. (1986). Intelligent information extraction from reflectance spectra: absorption band positions. /. Geophys. Res., 91, 9585-98. Hunt, G. (1980). Electromagnetic radiation: the communication link in remote sensing. In Remote Sensing in Geology, ed. B. Siegal and A. Gillespie (pp. 5-46). New York: Wiley. Hunt, G., and Vincent, R. (1968). The behavior of spectral features in the infrared emission from particulate surfaces of various grain sizes. /. Geophys. Res., 73, 6039-46. Irvine, W. (1965). Multiple scattering by large particles. Astrophys. J., 142, 1563-75. (1966). The shadowing effect in diffuse reflectance. /. Geophys. Res., 71, 2931-7. Irvine, W., and Pollack, J. (1968). Infrared properties of water and ice spheres. Icarus, 8, 324-60. Ishimaru, A. (1978). Wave Propagation and Scattering in Random Media. New York: Academic Press. Ishimaru, A., and Kuga, Y. (1982). Attenuation of a coherent field in a dense distribution of particles. /. Opt. Soc. Amer., 72, 1317-20. Jahnke, E., and Emde, E. (1945). Tables of Functions. New York: Dover Press. Jakowsky, B., Finiol, G., and Henderson, B. (1990). Directional variations in thermal emission from geologic surfaces. Geophys. Res. Lett., 17, 985-8. Jenkins, F., and White, H. (1950). Fundamentals of Optics, 2nd ed. New York: McGraw-Hill. Johnson, P., Smith, M., and Adams, J. (1985). Quantitative analysis of planetary reflectance spectra with principal components analysis. In Proc. 15th Lunar Planet. Sci. Conf, ed. G. Ryder and G. Schubert (pp. C805-10). Washington: American Geophysical Union. Johnson, P., Smith, M., Taylor-George, S., and Adams, J. (1983). A semiempirical method for analysis of the reflectance spectra of binary mineral mixtures. /. Geophys. Res., 88, 3557-61. Johnson, R., Nelson, M., McCord, T., and Gradie, J. (1988). Analysis of Voyager images of Europa: plasma bombardment. Icarus, 75, 423-36.
442
Bibliography
Joseph, J., Wiscombe, W., and Weinman, J. (1976). The delta-Eddington approximation for radiative flux transfer. /. Atmos. Sci., 33, 2452-9. Kattawar, G. (1975). A three parameter analytic phase function for multiple scattering calculations. /. Quant. Spectr. Rad. Trans., 15, 839-49. (1979). Radiative transfer in spherical shell atmospheres. III. Application to Venus. Icarus, 40, 60-6. Kattawar, G., and Eisner, M. (1970). Radiation from a homogeneous isothermal sphere. Appl. Opt., 9, 2685-90. Kattawar, G., and Humphreys, T. (1980). Electromagnetic scattering from two identical pseudospheres. In Light Scattering by Irregularly Shaped Particles, ed. D. Schuerman (pp. 177-90). New York: Plenum Press. Kemp, J. (1974). Circular polarization of planets. In Planets, Stars and Nebulae Studied with Photopolarimetry, ed. T. Gehrels (pp. 607-16). Tucson: University of Arizona Press. KenKnight, G., Rosenberg, D., and Wehner, G. (1967). Parameters of the optical properties of the lunar surface powder in relation to solar wind bombardment. /. Geophys. Res., 72, 3105-29. Kerker, M. (1969). The Scattering of Light. New York: Academic Press. Kimes, D., and Kerchner, J. (1982). Irradiance measurement errors due to the assumption of a Lambertian reference panel. Rem. Sens. Environ., 12, 141-9. Kittel, C. (1976). Introduction to Solid State Physics, 5th ed. New York: Wiley. Kocinski, J., and Wojtczak, L. (1978). Critical Scattering Theory, An Introduction. New York: Elsevier Scientific. Kolokolova, L. (1985). On the influence of the structure of atmosphereless bodies' surfaces to the polarimetric characteristics of reflected light. Solar System Res., 19, 165-73. (1990). Dependence of polarization on optical and structural properties of the surfaces of atmosphereless bodies. Icarus, 84, 305-14. Kortum, G. (1969). Reflectance Spectroscopy. New York: Springer-Verlag. Kourganoff, V. (1963). Basic Methods in Transfer Problems, Radiative Equilibrium and Neutron Diffusion. New York: Dover Press. Kubelka, P. (1948). New contributions to the optics of intensely light-scattering materials. Part I. /. Opt. Soc. Amer., 38, 448-57. (1954). New contributions to the optics of intensely light-scattering materials. Part II. Nonhomogeneous layers. /. Opt. Soc. Amer., 44, 330-5. Kubelka, P., and Munk, F. (1931). Ein Beitrag zur Optik der Farberntricke. Zeitschr. Techn. Physik, 12, 593-601. Kuga, Y., and Ishimaru, A. (1984). Retroreflection from a dense distribution of spherical particles. /. Opt. Soc. Amer., 8, 831-5. Landau, L., and Lifschitz, E. (1975). The Classical Theory of Fields, 4th ed. New York: Pergamon Press. Lass, H. (1950). Vector and Tensor Analysis. New York: McGraw-Hill.
Bibliography
443
Lax, M. (1954). The influence of lattice vibrations on electronic transitions in solids. In Photoconductivity Conference, ed. R. Breckenridge, B. Russell, and E. Hahn (pp. 111-45). New York: Wiley. Leinert, C, Link, H., Pitz, E., and Giese, R. (1976). Interpretation of a rocket photometry of the inner zodiacal light. Astron. Astrophys., 47, 221-30. Lenoble, J. (1985). Radiative Transfer in Scattering and Absorbing Atmospheres. Hampton, Va.: Deepak Publishing. Liou, K., Cai, Q., and Pollack, J. (1983). Light scattering by randomly oriented cubes and parallelepipeds. Appl. Opt, 22, 3001-8. Liou, K., and Coleman, R. (1980). Light scattering by hexagonal columns and plates. In Light Scattering by Irregularly Shaped Particles, ed. D. Schuerman (pp. 207-18). New York: Plenum Press. Liou, K., and Hansen, J. (1971). Intensity and polarization for single scattering by polydisperse spheres: a comparison of ray optics and Mie theory. /. Atmos. ScL, 28, 995-1004. Liou, K., and Schotland, R. (1971). Multiple backscattering and depolarization from water clouds for a pulsed lidar system. /. Atmos. Set, 28, 772-84. Logan, L., Hunt, G., Salisbury, J., and Balsamo, S. (1973). Compositional implications of Christiansen frequency maximums for infrared remote sensing applications. /. Geophys. Res., 78, 4983-5003. Lorentz, H. (1952). The Theory of Electrons. New York: Dover Press. Lumme, K., and Bowell, E. (1981). Radiative transfer in the surfaces of atmosphereless bodies. I. Theory. Astron. J., 86, 1694-704. Lyot, B. (1929). Recherches sur la polarisation de la lumiere des planetes et de quelques substances terrestres. Ann. Obs. Paris, vol. VIII, Book 1 (translated as NASA Tech. Transl. TT-F-187, 1964). McEwan, A. (1991). Photometric functions for photoclinometry and other applications. Icarus, 92, 298-311. McGuire, A. (1993). Experimental Investigation of Light Scattering by Large, Irregular Particles. Ph.D. dissertation, University of Pittsburgh. McKay, D., Fruland, R., and Heiken, G. (1974). Grain size and the evolution of lunar soils. In Proc. 5th Lunar Sci. Conf, ed. W. Gose (pp. 887^906). New York: Pergamon Press. MacKintosh, F., and John, S. (1988). Coherent backscattering of light in the presence of time-reversal, non-invariant and parity-violating media. Phys. Rev., B37, 1884-97. Margenau, H., and Murphy, G. (1956). The Mathematics of Physics and Chemistry. New York: Van Nostrand. Marion, J. (1965). Classical Electromagnetic Radiation. New York: Macmillan. Matson, D., and Brown, R. (1988). Solid state greenhouses and their implications for icy satellites. Icarus, 77, 67-81. Maxwell-Garnett, J. (1904). Colours in metal glasses and in metallic films. Phil. Trans. Roy. Soc, A203, 385-420.
444
Bibliography
Melamed, N. (1963). Optical properties of powders. Part 1. Optical absorption coefficients and the absolute value of the diffuse reflectance. Part II. Properties of luminescent powders. /. Appl. Phys., 34, 560-70. Middleton, W., and Sanders, C. (1951). The absolute spectral diffuse reflectance of magnesium oxide. /. Opt. Soc. Amer., 41, 419-24. Minnaert, M. (1941). The reciprocity principle in lunar photometry. Astrophys. J., 93, 403-10. (1961). Photometry of the moon. In Planets and Satellites, ed. G. Kuiper and B. Middlehurst (pp. 213-45). Chicago: University of Chicago Press. Montgomery, W., and Kohl, R. (1980). Opposition effect experimentation. Opt. Lett., 5, 546-8. Morris, R., Lauer, H., Lawson, C, Gibson, E., Nace, G., and Stewart, C. (1985). Spectral and other physicochemical properties of submicron powders of hematite, maghemite, magnetite, goethite and lepidocrocite. /. Geophys. Res., 90, 3126-44. Morse, P., and Feshbach, H. (1953). Methods of Theoretical Physics. New York: McGraw-Hill. Muhleman, D. (1964). Radar scattering from Venus and the moon. Astron. I, 69, 34-41. Muinonen, K. (1990). Light scattering by inhomogeneous media: backward enhancement and reversal of linear polarization. Ph.D. dissertation. University of Helsinki, Finland. Muinonen, K., Lumme, K., Peltoniemi, J., and Irvine, W. (1989). Light scattering by randomly oriented crystals. Appl. Opt, 28, 3051-60. Mukai, S., Mukai, T., Giese, R., Weiss, K., and Zerull, R. (1982). Scattering of radiation by a large particle with a random rough surface. Moon and Planets, 26, 197-208. Mustard, J., and Pieters, C. (1987). Quantitative abundance estimates from bidirectional reflectance measurements. In Proc. 17th Lunar Planet. Sci. Conf., ed. G. Ryder and G. Schubert (pp. E617-26). Washington: American Geophysical Union. (1989). Photometric phase functions of common geologic minerals and applications to quantitative analysis of mineral mixture reflectance spectra. /. Geophys. Res., 94, 13619-34. Nash, D. (1983). Io's 4-/x,m band and the role of adsorbed SO2. Icarus, 54, 511-23. (1986). Mid-infrared reflectance spectra (2.3-22 fim) of sulfur, gold, KBr, MgO and halon. Appl. Opt., 25, 2427-33. Nash, D., and Conel, J. (1974). Spectral reflectance systematics for mixtures of powdered hypersthene, labradorite and ilmenite. /. Geophys. Res., 79, 1615-21. Nicklasson, G., Granqvist, C, and Henderi, O. (1981). Effective medium models for the optical properties of inhomogeneous materials. Appl. Opt, 20, 26-30.
Bibliography
445
Nicodemus, F. (1970). Reflectance nomenclature and directional reflectance and emissivity. Appl. Opt.y 9, 1474-5. Nicodemus, R, Richmond, J., Hsia, J., Ginsberg, I., and Limperis, T. (1977). Geometrical Considerations and Nomenclature for Reflectance. National Bureau of Standards Monograph 160. Nitsan, U., and Shankland, T. (1976). Optical properties and electronic structure of mantle silicates. Geophys. J. Roy. Astron. Soc, 45, 59-87. O'Donnell, K., and Mendez, E. (1987). Experimental study of scattering from characterized random surfaces. /. Opt. Soc. Amer.y A4, 1194-205. Oetking, P. (1966). Photometric studies of diffusely reflecting surfaces with applications to the brightness of the moon. /. Geophys. Res., 71, 2505-13. Ohman, Y. (1955). A tentative explanation of the negative polarization in diffuse reflection. Ann. Obs. Stockholm, 18(8), 1-10. Ostro, S. (1982). Radar properties of Europa, Ganymede and Callisto. In Satellites of Jupiter, ed. D. Morrison (pp. 213-36). Tucson: University of Arizona Press. Ostro, S., and Shoemaker, E. (1990). The extraordinary radar echoes from Europa, Ganymede and Callisto: a geological perspective. Icarus, 85, 335-45. Otterman, J. (1983). Absorption of insolation by land surfaces with sparse vertical protrusions. Tellus, 35b, 309-18. Palik, E. (ed.) (1991). Handbook of Optical Constants of Solids. New York: Academic Press. Paliouras, J. (1975). Complex Variables for Scientists and Engineers. New York: Macmillan. Panofsky, W., and Phillips, M. (1962). Classical Electricity and Magnetism. Cambridge, MA: Addison-Wesley. Perrin, J., and Lamy, P. (1983). Light scattering by large rough particles. Optica Acta, 30, 1223-44. Perry, R., Hunt, A., and Huffman, D. (1978). Experimental determinations of Mueller scattering matrices for non-spherical particles. Appl. Opt., 17, 2700-10. Pinnick, R., Carroll, D., and Hofmann, D. (1976). Polarized light from monodisperse randomly oriented nonspherical aerosol particles: measurements. Appl. Opt., 15, 384-93. Pinty, B., and Verstraete, M. (1991). Extracting information on surface properties from bidirectional reflectance measurements. /. Geophys. Res., 96, 2865-74. Pinty, B., Verstraete, M., and Dickinson, R. (1989). A physical model for predicting bidirectional reflectances over bare soil. Rem. Sens. Environ., 27, 273-88. (1990). A physical model of the bidirectional reflectance of vegetation canopies. 2. Inversion and validation. /. Geophys. Res., 95, 11767-75.
446
Bibliography
Pollack, J., and Cuzzi, J. (1980). Scattering by nonspherical particles of size comparable to a wavelength: a new semi-empirical theory and its application to tropospheric aerosols. /. Atmos. Sci., 37, 868-81. Pollack, J., and Whitehill, L. (1972). A multiple scattering model of the diffuse component of the lunar radar echoes. /. Geophys. Res., 77, 4289-303. Rayleigh, Lord (1871). On the light from the sky, its polarization and colour. Philos. Mag., 41, 107-20, 274-9. Reichman, J. (1973). Determination of absorption and scattering coefficients for nonhomogeneous media. I. Theory. Appl. Opt., 12, 1811-23. Richter, N. (1962). The photometric properties of interplanetary matter. Quart. J. Roy. Astron. Soc, 3, 179-86. Ross, J., and Marshak, A. (1984). Calculation of the canopy bidirectional reflectance using the Monte-Carlo method. Rem. Sens. Environ., 24, 213-25. Rozenberg, G. (1966). Twilight. New York: Plenum Press. Russell, H. (1916). On the albedo of planets and their satellites. Atrophys. J., 43, 173-87. Salisbury, J. (1993). Mid-infrared spectroscopy: laboratory data. In Remote Geochemical Analysis, ed. C. Pieters and P. Englert. Cambridge University Press. Salisbury, J., and Eastes, J. (1985). The effect of particle size and porosity on spectral contrast in the mid-infrared. Icarus, 64, 586-8. Salisbury, J., Hapke, B., and Eastes, J. (1987). Usefulness of weak bands in midinfrared remote sensing of particulate planetary surfaces. /. Geophys. Res., 92, 702-10. Salisbury, J., and Wald, A. (1992). The role of volume scattering in reducing spectral contrast of restrahlen bands in spectra of powdered minerals. Icarus, 96, 121-8. Salisbury, J., and Walter, L. (1989). Thermal infrared (2.5-13.5 ju,m) spectroscopic remote sensing of igneous rock types on particulate planetary surfaces. /. Geophys. Res., 94, 9192-202. Saunders, P. (1967). Shadowing on the ocean and the existence of the horizon. /. Geophys. Res., 72, 4643-9. Schaber, G., Berlin, G., and Brown, W., Jr. (1976). Variations in surface roughness within Death Valley, California: geologic evaluation of 25-cm wavelength radar images. Geol. Soc. Am. Bull., 87, 29-41. Schatz, E. (1966). Effect of pressure on the reflectance of compacted powders. /. Opt. Soc. Amer., 56, 389-94. Schiffer, R., and Thielheim, K. (1982a). Light reflection from randomly oriented convex particles with rough surfaces. /. Appl. Phys., 53, 2825-30. (1982b). A scattering model for the zodiacal light particles. Astron. Astrophys., 116, 1-9.
Bibliography
447
Schlatter, T. (1972). The local surface energy balance and subsurface temperature regime in Antarctica. /. Appl. Meteor., 11, 1048-62. Schonberg, E. (1929). Theoretische Photometrie. In Handbuch der Astrophysik, Vol. 2, ed. G. Eberhard, A. Kohlschutter, and H. Ludendorff (pp. 1-280). Berlin: Julius Springer. Schuerman, D. (1980). Light Scattering by Irregularly Shaped Particles. New
York: Plenum Press. Schuerman, D., Wang, R., Gustafson, B., and Schaefer, R. (1981). Systematic studies of light scattering. 1: Particle shape. Appl Opt., 20, 4039-50. Schulman, J., and Compton, W. (1962). Color Centers in Solids. New York: Pergamon Press. Schuster, A. (1905). Radiation through a foggy atmosphere. Astrophys. J., 21, 1-22. Seeliger, H. (1887). Zur Theorie der Beleuchtung der grossen Planeten inbesondere des Saturn. Abhandl Bayer. Akad. Wiss. Math.-Naturw. Kl. II, 16, 405-516. (1895). Theorie der Beleuchtung staubformiger kosmischen Masses insbesondere des Saturinges. Abhandl. Bayer. Akad. Wiss. Math.-Naturw. Kl. II, 18, 1-72. Shkuratov, Y. (1982). A model for negative polarization of light by cosmic bodies without atmospheres. Sou. Astron., 26, 493-6. (1988). Diffractional model of the brightness surge of complex structure surfaces. Kin., Phys., Cel. Bodies, 4, 33-9. (1989). New mechanism of formation of negative polarization of light scattered by the solid surfaces of cosmic bodies. Solar System Res., 23, 111-13. Simonelli, D., and Veverka, J. (1987). Phase curves of minerals on Io: interpretation in terms of Hapke's function. Icarus, 68, 503-21. Simpson, R., and Tyler, G. (1982). Radar scattering laws for the lunar surface. IEEE Trans. Antennas and Propagation, AP30, 438-49. Skorobogatov, B., and Usoskin, A. (1982). Optical properties of ground surfaces of nonabsorbing materials. Opt. Spectr., 52, 310-13. Smith, D. (1985). Dispersion theory, sum rules and their application to the analysis of optical data. In Handbook of Optical Constants of Solids, ed. E. Palik (pp. 35-154). New York: Academic Press. Smith, J. (1983). Matter-energy interactions in the optical region. In Manual of Remote Sensing, 2nd ed., ed. R. Colwell (pp. 61-113). Falls Church, VA: American Society of Photogrammetry. Smith, M., Johnson, P., and Adams, J. (1985). Quantitative determination of mineral types and abundances from reflectance spectra using principal components analysis. In Proc. 15th Lunar Planet. Sci. Conf, ed. G. Ryder and G. Schubert (pp. C797-804). Washington: American Geophysical Union. Smythe, W. (1975). Spectra of hydrate frosts: their application to the outer solar system. Icarus, 24, 421-7.
448
Bibliography
Sobolev, V. (1975). Light Scattering in Planetary Atmospheres. New York: Pergamon Press. Sokolov, A. (1967). Optical Properties of Metals. New York: Elsevier. Spencer, J. (1990). A rough-surface thermophysical model for airless planets. Icarus, 83, 27-38. Spitzer, W., and Kleinman, D. (1961). Infrared lattice bands of quartz. Phys. Rev., 121, 1324-35. Sproull, R., and Phillips, W. (1980). Modern Physics, 3rd ed. New York: Wiley. Steigman, G. (1978). A polarimetric model for a dust covered planetary surface. Mon. Not. Roy. Astron. Soc, 185, 877-88. Stratton, J. (1941). Electromagnetic Theory. New York: McGraw-Hill. Stroud, D., and Pan, F. (1978). Self-consistent approach to electromagnetic wave propagation in composite media: application to model granular metals. Phys. Rev., B17, 1602-10. Suits, G. (1972). The calculation of the directional reflectance of a vegetative canopy. Rem. Sens. Environ., 2, 117-25. Sung, C, Singer, R., Parkin, K., and Burns, R. (1977). Temperature dependence of Fe 2+ crystal field spectra: implications to mineralogical mapping of planetary surfaces. In Proc. 8th Lunar Sci. Conf., ed. R. Merrill (pp. 1063-79). New York: Pergamon Press. Tanashchuk, M., and Gilchuk, L. (1978). Experimental scattering matrices of ground glass surfaces. Opt. Spectr., 45, 658-62. Thompson, T., Pollack, J., Campbell, M., and O'Leary, B. (1970). Radar maps of the moon at 70 cm wavelength and their interpretation. Rad. Sci., 5, 253-62. Thorpe, T. (1973). Mariner 9 photometric observations of Mars from November 1971 through March 1972. Icarus, 20, 482-9. (1978). Viking orbiter observations of the Mars opposition effect. Icarus, 36, 204-15. Torrance, K., and Sparrow, E. (1967). Theory for off-specular reflection from roughened surfaces. /. Opt. Soc. Amer., 57, 1105-14. Trowbridge, T. (1978). Retroreflection from rough surfaces. /. Opt. Soc. Amer., 68, 1225-42. Umov, N. (1905). Chromatische Depolarization durch Lichtzerstreuung. Phys. Z., 6, 674-6. Ungut, A., Grehan, G., and Gouesbet, G. (1981). Comparisons between geometrical optics and Lorentz-Mie theory. Appl. Opt., 20, 2911-18. Van de Hulst, H. (1957). Light Scattering by Small Particles. New York: Wiley. (1974). The spherical albedo of a planet covered with a homogeneous cloud layer. Astron. Astrophys., 35, 209-14. (1980). Multiple Light Scattering. New York: Academic Press.
Bibliography
449
Vanderbilt, V., Grant, L., Biehl, L., and Robinson, B. (1985). Specular, diffuse and polarized light scattered by two wheat canopies. Appl. Opt., 24, 2408-18. Van Diggelen, J. (1959). Photometric properties of lunar crater floors. Rech. Obs. Utrecht, 14, 1-114. (1965). The radiance of lunar objects near opposition. Planet. Space Sci., 13, 271-9. Vaughan, D. (1990). Some contributions of spectral studies in the visible and near visible light region to mineralogy. In Absorption Spectroscopy in Mineralogy, ed. A. Mottana and T. Burragato (pp. 1-37). New York: Elsevier. Verstraete, M., Pinty, B., and Dickinson, R. (1990). A physical model of the bidirectional reflectance of vegetation canopies. 1. Theory. /. Geophys. Res., 95, 11755-65. Veverka. J., Goguen, J., Yang, S., and Elliot, J. (1978a). Near-opposition limb darkening of solids of planetary interest. Icarus, 33, 368-79. (1978b). How to compare the surface of Io to laboratory samples. Icarus, 34, 63-7. (1978c). Scattering of light from particulate surfaces. 1. A laboratory assessment of multiple scattering effects. Icarus, 34, 406-14. Veverka, J., Helfenstein, P., Hapke, B., and Goguen, J. (1988). Photometry and polarimetry of Mercury. In Mercury, ed. F. Vilas and C. Chapman (pp. 37-58). Tucson: University of Arizona Press. Vincent, R., and Hunt, G. (1968). Infrared reflectance from mat surfaces. Appl Opt., 7, 53-9. Wagner, R. (1967). Shadowing of randomly rough surfaces. /. Acoust. Soc. Amer., 41, 138-47. Wallach, D., and Hapke, B. (1985). Light scattering in a spherical exponential atmosphere, with applications to Venus. Icarus, 63, 354-73. Walter, L., and Salisbury, J. (1989). Spectral characterization of igneous rocks in the 8 to 12 /im region. /. Geophys. Res., 94, 9203-13. Warren, S. (1982). Optical properties of snow. Rev. Geophys. Space Phys., 20, 67-89. Watson, G. (1958). A Treatise on the Theory of Bessel Functions. Cambridge University Press. Weidner, V., and Hsia, J. (1981). Reflection properties of pressed polytetrafluoroethylene powder. /. Opt. Soc. Amer., 71, 856-61. Weidner, V., Hsia, J., and Adams, B. (1985). Laboratory intercomparison study of pressed polytetrafluoroethylene powder reflectance standards. Appl. Opt, 24, 2225-30. Weiss-Wrana, K. (1983). Optical properties of interplanetary dust: comparison with light scattering by larger meteoritic and terrestrial grains. Astron. Astrophys., 126, 240-50. Wells, E. (1977). Optical Absorption Bands in Glasses of Lunar Composition. Ph.D. dissertation, University of Pittsburgh.
450
Bibliography
Wells, E., and Hapke, B. (1977). Lunar soil: iron and titanium bands in the glass fraction. Science, 195, 977-9. Wells, E., Veverka, J., and Thomas, P. (1984). Mars: experimental study of albedo changes caused by dust fallout. Icarus, 58, 331-8. Wendtland, W., and Hecht, H. (1966). Reflectance Spectroscopy. New York: Wiley-Interscience. Whitaker, E. (1969). An investigation of the lunar heiligenschein. In Analysis of Apollo 8 Photography and Visual Observations (pp. 38-9). NASA SP-201. White, W., and Keester, K. (1966). Optical absorption spectra of iron in the rock-forming silicates. Amer. Min., 51, 774-91. Widorn, T. (1967). Zur photometrischen Bestimmung der Durchmesser der kleinen Planeten. Ann. Univ. Sternw. Wien, 27, 112-19. Wildey, R. (1978). The moon in heiligenschein. Science, 200, 1265-7. Woessner, P., and Hapke, B. (1987). Polarization of light scattered by clover. Rent. Sens. Environ., 21, 243-61. Wolf, P., and Maret, G. (1985). Weak localization and coherent backscattering of photons in disordered media. Phys. Rev. Lett., 55, 2696-9. Wolff, M. (1975). Polarization of light reflected from rough planetary surface. Appl. Opt, 14, 1395-405. (1980). Theory and application of the polarization-albedo rules. Icarus, 44, 780-92. (1981). Computing diffuse reflection from particulate planetary surface with a new function. Appl Opt., 20, 2493-8. Wooten, F. (1972). Optical Properties of Solids. New York: Academic Press. Yolken, H., and Kruger, J. (1965). Optical constants of iron in the visible region. /. Opt. Soc. Amer., 55, 842-4. Young, A. (1973). Are the clouds of Venus sulfuric acid? Icarus, 18, 564-82. Zellner, B., Gehrels, T., and Gradie, J. (1974). Polarimetric diameters. Astron. J.y 79, 1100-10. Zellner, B., and Gradie, J. (1976). Polarimetric evidence for the albedos and compositions of 94 asteroids. Astron. J., 81, 262-80. Zerull, R. (1976). Scattering measurements of dielectric and absorbing nonspherical particles. Contr. Atmos. Phys., 49, 168-88. Zerull, R., and Giese, R. (1974). Microwave analogue studies. In Planets, Stars and Nebulae Studied with Photopolarimetry, ed. T. Gehrels (pp. 901-15). Tucson: University of Arizona Press.
INDEX
absorbance, 300 absorption band shape, 42 bands in reflectance, 303 coefficient, 23, 319 length, 23 of light, 24 mechanisms, 31 in reflectivity, 59 volume coefficient, 152, 157, 172 albedo, 271 bolometric, 278 Bond, see reflectance, spherical factor, 173 geometric, 273, 287 hemispherical, 183, 273 normal, 272, 288 physical, 273, 287 plane, see reflectance, hemispherical single-scattering: calculating from reflectance, 290; of mixture, volume, 156, 280; of particle, 66; of particulate medium, average, 172 spherical, see reflectance, spherical Allen diffraction function, 109, 146 angular scattering phase function: of mixture, average, 280; particle, 67; volume, 156, 173 volume coefficient, 153, 158, 172 associated Legendre functions, 423 asteroids, 403 asymmetry factor cosine, 68 hemispherical, 68
Babinet's principle, 83 band, model of electrons in solid, 35 conduction, 35 transitions between, 40 valence, 35 Becke method, 293 Bessel functions, 425 bidirectional-reflectance distribution function, 261 black-body radiation, 360 boundary conditions, 43 Brewster's angle, 54, 293, 395 bright shadow, see opposition effect brightness, 64 temperature, 379 Cauchy-Gorsat theorem, 418 Cauchy integral formula, 418 Cauchy-Riemann condition, 418 Chandrasekhar, S., 4 charge, electric, 6 charge-transfer band, 39 Christiansen frequency, 59, 384 Clausius-Mossotti relation, 14, 148 cloudbow, 92, 117 coherent backscatter, 229, 410 coherent effects, 170 cohesion between particles, 160 collision frequency, 25 color center, 36 complex variables, 413 conductivity electric, 19, 26 optical, 21 constitutive equations, 7 critical opalescence, 150
451
452
Index
cross section absorption, 66, 102 extinction, 65, 102 geometric, 66, 102 scattering, 66, 102 crossing symmetry, 419 crystal-field theory, 37 current, electric, 6 delta function, 64 dielectric constant, 13, 20 of ice, 33 of water, 33 diffraction, 78 of disk, 83 of hole, 78 of irregular particle, 123 of isolated sphere, 83 of particles that are not isolated, 86, 162 dilution method, 321, 322 dipoles electric, 12, 148 magnetic, 16 scattering by, 103 dispersion, anomalous, 28 normal, 28 quantum-mechanical theory, 30 relation, 23, 30 displacement, electric, 6, 13, 148 doubling method, 175 Drude model, 24 dust, lunar, xii Eddington approximation, 178 effective-medium theory, 147, 302 effective particle density, 169 effective particle size, 98, 130, 297 effective single-particle absorption thickness, see espat function efficiency absorption, 66, 102 extinction, 66, 102 scattering, 66, 102, 123 eikonal approximation, 115 electronic transitions, 35 embedded invariance method, 180, 205, 365 emergence, plane of, 185 emission and reflectance combined, 375 thermal, 154, 358 volume coefficient, 154 emissivity, 362 directional, 365, 374 effect of particle size on, 380 factor, 364, 374
hemispherical, 372 integrated, 362 spectral, 362 emittance, 362 error function, 224, 258 approximation for, 258 complementary, 258 espat function, 66, 139, 298, 322 Europa, 351, 357 exciton, 39 extended-boundary-condition method, 115 extinction cross-section, 65, 102 efficiency, 66, 102 volume coefficient, 152, 156, 172 fairy castle structures, xiii, 218 field electric, 6 magnetic induction, 6 filling factor, 149, 163 free carriers, 41 Fresnel equations, 44 Galileo, 1 Gaussian quadratures method, 178 Gauss's theorem, 416 geometric-optics scattering, 77, 86 glory, 92, 117 Gold, T., xii gold black, 303 gradient, vector, 414 gray body, 362 H functions, 209 approximations for, 212 moments of, 210 Hankel functions, 425 heilegenschein, see opposition effect Helmholtz's equation, 423 Henyey-Greenstein function, 108 hole, 36 hot spot, see opposition effect ice, 33, 306 incidence, plane of, 185 integral phase function, 275 intensity, 64 magnetic, 6 specific, 64 internal scatterers, 116, 131, 138 internal-transmission factor, 98, 129 irradiance, 11
Index KirchhofFs law, 363, 373 Kramers-Kronig relations, 31, 58 Kubelka-Munk theory, 4, 183, 287, 318 Lambert sphere, 110 Lambert surface, 191 Lambert's law, 111, 190, 197, 235 Laplace's equation, 423 Laporte rule, 38 lattice vibrations, 33 acoustical branch, 34 optical branch, 34 layers effect of, on absorption bands, 314 reflectance of, 243, 251 leaf-area index, 243 leaf-droop index, 241 Legendre polynomials, 107, 424 Lommel-Seeliger law, 111, 199 Lommel-Seeliger sphere, 110 Lorentz-Lorenz relations, see Clausius-Mossotti relation Lorentz model, 26 loss tangent, 20 luminance coordinates, 186 magnetization, 16 Maxwell-Garnett model, 148 fallacy of, 149 role of density fluctuations in, 149 Maxwell's field equations, 6 mean optical path length, 300 Melamed model, 130 Mercury, 351, 357 Mie theory, 69 Minnaert albedo, 192 Minnaert index, 192 Minnaert's law, 191 mixtures, 278 areal, 278 espat of, 283 intimate, 279 spectra of, 310 molecular rotation, 32 Monte Carlo method, 174 moon, 1, 356, 386 Mueller matrix, 388 multistream method, 177 Murphy's law, 5 Ohm's law, 19 opposition effect, 169, 216 amplitude of, 228 amplitude of mixture, 281 angular width of, 225
453
angular width of mixture, 281 of vegetation canopies, 242 optical coupling, 169 optical flatness, Rayleigh criterion for, 61, 125 optical path, 300 orthogonal set, 427 oscillator strength, 30 permeability of free space, 7 magnetic, 17 permittivity electric, 13 of free space, 7 phase, 8 angle, 66, 185 function of particulate medium, volume, 173 integral, 278 single-particle function, 67, 102 phonons, 33, 41 photometric function, 272 photon traps, 303 Planck function, 360 plasma frequency, 25 plasma resonance, 74, 125 polar darkening, 351 polarizability, electric, 13, 25, 148 polarization -albedo relation, 400 and Brewster's angle, 395 circular, 387, 389 effect of transmitted component, 397 electric, 13, 148 elliptical, 58, 387 from irregular particles, 118 linear, 3, 35, 386 negative, 389 phase curve, 390 positive, 389 of sphere, 96 of vegetation, 398 porosity, 149 Poynting vector, 11 principal plane, 185 radar, 230, 283 radialith, 317 radiance, 64 radiative transfer, equation of, 151 in medium of well-separated scatterers, 157 in particulate medium of arbitrary separations, 159, 171 problems, methods of solution, 174 radiosity method, 175 radius, equivalent of irregular particle, 102
454
Index
rainbow, 92 Rayleigh absorber, 74, 125 region, 73 scatterer, 74 Rayleigh-Jeans law, 361 reciprocity principle, 263, 346 reflectance and absorption coefficient, 292 biconical, 183 bidirectional, 181, 223, 286, 291 bihemispherical, 183,184, 269, 277, 287, 292 coefficient, 262 combined specular and diffuse, 236 diffuse, 191 diffusive, 4, 183, 192, 269, 291 directional-hemispherical, 182,183,184,265, 286, 291 effect of particle size on, 292 effect of thermal emission on, 382 factor, 262 hemispherical, 182, 183, 184, 265, 286, 291 hemispherical-directional, 182,183,268,286, 292 inversion of, 288 Lambert, 190 of layered media, 243, 251, 314 measurement of, 285 Minnaert, 190 of mixtures, 278, 310 relative, 262, 291 roughness, effects of, 325 spectroscopy, 284 spherical, 183, 184, 269, 277, 287, 288, 292 in spherical geometry, 255 reflection, normal, 46 of dielectric, 52 from exterior surface of particle, 87, 124 from interior surface of particle, 128 of metals, 48 regular, 236 specular, 43 refraction, 49 index of, 214 in sphere, 89 regolith, lunar, 120, 226, 235, 401 relaxation time, 32 remission function, 320 restrahlen, 60 roughness effects of, on planetary photometry, 345 effects of, on reflectance, 325 Russell's rule, 278
scattering angle, 66 by colloids, 105* 160 by cube, 104, 114 by cylinder, 100, 104 by irregular particle, 101 by oriented particles, 239 plane, 65, 85 by sphere, 63 by spheroid, 104 volume coefficient, 153, 158, 172 Schonberg function, see Lambert sphere Schuster-Schwarzschild method, see two-stream method separation-of-variables method, 420 shadowing function, 331 interparticle, 169, 216 large scale roughness, 331 shadows projected, 333 tilt, 333 similarity relations, 197, 270 size parameter, 71 slope -albedo relation in polarization, 402 distribution function, 326 mean angle of, 326 Snell's law, 49 source function, 156, 364 specific inductive capacity, 13 spectroscopy emittance, 379 reflectance, 284 spherical harmonics, 426 spin-multiplicity rule, 38 Stefan-Boltzmann law, 362 Stokes theorem, 416 Stokes vector, 387 successive-approximation method, 179 sum rule, 29 surface-scattering region strong, 297 weak, 297 susceptibility electric, 13, 25 magnetic, 17 thermal black-body radiation, 360 emission, 358 infrared, 358 source function, 364 volume emission coefficient, 364 thermodynamic enclosure, 361
Index
Umov effect, 400
vegetation polarization by, 398 scattering by, 239 Venus, 114, 260, 387 volume absorption coefficient, 152, 157, 172 emission coefficient, 154 extinction coefficient, 152, 156, 172 scattering coefficient, 153, 158, 172 -scattering region, 295
vector calculus, 413 cross-product, 413 divergence of, 414 dot product, 413 gradient of, 414
water, 33 wave equation, 7, 420 weak photon localization, see coherent backscatter Wien displacement law, 361
tilt, effective angles of, 331 transition minimum, 306 transmission of dielectric, 52 normal, 46 of slab of particles, 44 specular, 43 transparency feature, 384 two-stream method, 178, 193, 201
455