Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation...

20
ELSEVIER Optical Remote Sensing of Vegetation: Modeling, Caveats, and Algorithms R. B. Myneni, B. Pinty, ** D. * S. Maggion, S. Kimes, * ** J. Iaquinta, J. L. Privette,t N. Gobron, 1. M. Verstraete,tt and D. L. Williams* The state-of-the-art on radiative transfer modeling in vegetation canopies and the application of such models to the interpretation and analysis of remotely sensed optical data is summarized. Modeling of top-of-the-atmo- sphere and top-of-the-canopy radiance field is developed as boundary value problems in radiative transfer. The parameterization of the constituentfunctions with simple models and/or empirical data is outlined togetherwith numerical solution methods and examples of results of model validation. Caveats in the assignment of signal characteristics to surface properties are itemized and discussed with example results. Algorithms to estimate surfaceproperties from remote observations are classified as spectral vegetation indices, model inversion, expert systems, neural networks, and genetic algorithms. Their applicability is also discussed. I. INTRODUCTION The interpretation of remotely sensed reflectance data and the apportioning of signal characteristics to target properties requires an understanding of how the various physical mechanisms interact to produce the measured signal. For instance, in the application of satellite remote sensing to vegetation, an understanding of the spectral response resulting from leaf internal microscopic struc- ture and the perturbations to this response introduced by both the macroscopic aggregation of leaves in a Biospheric Sciences Branch, NASA-Goddard Space Flight Center, Greenbelt, MD Universit6 Blaise Pascal, Laboratoire de Mteorologie Phy- sique, Aubiiere, France tDepartment of Aerospace Engineering Studies, University of Colorado, Boulder ttInstitute for Remote Sensing Applications, CEC Joint Research Center, Ispra (VA), Italy Address correspondence to R. B. Myneni, Goddard Space Flight Center, Code 923, Greenbelt, MD 20771. R. B. Myneni is a University of Maryland Association Resident Associate. Received 2 August 1993; accepted 8 August 1994. REMOTE SENS. ENVIRON. 51:169-188 (1995) ©Elsevier Science Inc., 1995 655 Avenue of the Americas, New York, NY 10010 canopy and the intervening atmosphere is required. This physical problem may be succinctly posed as a radiative transfer equation, the solution of which con- volved with an instrument's response function is the remote spectral measurement. In this paper, the state-of-the-art on modeling of radiation transport in vegetation canopies and the appli- cation of such models to the interpretation and analysis of remotely sensed optical data is summarized. We begin with a statement of the physical problem and its mathematical description, parameterization and numeri- cal solution (section II). This is followed by a brief catalogue of the various distortionary effects inherent in remote observations (section III). The extraction of information (namely estimation of vegetation properties) from remote optical measurements is discussed in the last section. This paper is not intended to be a review either of literature or of the various ideas currently in vogue. Instead, it is aimed with the broader scope of itemizing and elaborating three issues in the optical remote sensing of vegetation-modeling, caveats, and algorithms. II. THE PHYSICAL PROBLEM In this section we begin with a brief statement of the physical problem encountered in the optical remote sensing of vegetated land surfaces. Forward modeling of the physical problem is developed as boundary value problems in radiative transfer. The parameterization and solution of the resulting equations are discussed along with results of validation of the models. Statement of the Physical Problem The atmosphere is illuminated at the top by monodirec- tional solar radiation that is absorbed and / or scattered depending on the composition of the atmosphere and wavelength of the incident beam. The vegetation canopy 0034-4257 / 95 / $9.50 SSDI 0034-4257(94)00073-V **

Transcript of Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation...

Page 1: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

ELSEVIER

Optical Remote Sensing of Vegetation:Modeling, Caveats, and Algorithms

R. B. Myneni,B. Pinty, ** D.

* S. Maggion,S. Kimes, *

** J. Iaquinta, J. L. Privette,t N. Gobron,1. M. Verstraete,tt and D. L. Williams*

The state-of-the-art on radiative transfer modeling invegetation canopies and the application of such modelsto the interpretation and analysis of remotely sensedoptical data is summarized. Modeling of top-of-the-atmo-sphere and top-of-the-canopy radiance field is developedas boundary value problems in radiative transfer. Theparameterization of the constituentfunctions with simplemodels and/or empirical data is outlined together withnumerical solution methods and examples of results ofmodel validation. Caveats in the assignment of signalcharacteristics to surface properties are itemized anddiscussed with example results. Algorithms to estimatesurface properties from remote observations are classifiedas spectral vegetation indices, model inversion, expertsystems, neural networks, and genetic algorithms. Theirapplicability is also discussed.

I. INTRODUCTION

The interpretation of remotely sensed reflectance dataand the apportioning of signal characteristics to targetproperties requires an understanding of how the variousphysical mechanisms interact to produce the measuredsignal. For instance, in the application of satellite remotesensing to vegetation, an understanding of the spectralresponse resulting from leaf internal microscopic struc-ture and the perturbations to this response introducedby both the macroscopic aggregation of leaves in a

Biospheric Sciences Branch, NASA-Goddard Space FlightCenter, Greenbelt, MD

Universit6 Blaise Pascal, Laboratoire de Mteorologie Phy-sique, Aubiiere, France

tDepartment of Aerospace Engineering Studies, University ofColorado, Boulder

ttInstitute for Remote Sensing Applications, CEC Joint ResearchCenter, Ispra (VA), Italy

Address correspondence to R. B. Myneni, Goddard Space FlightCenter, Code 923, Greenbelt, MD 20771. R. B. Myneni is a Universityof Maryland Association Resident Associate.

Received 2 August 1993; accepted 8 August 1994.

REMOTE SENS. ENVIRON. 51:169-188 (1995)©Elsevier Science Inc., 1995

655 Avenue of the Americas, New York, NY 10010

canopy and the intervening atmosphere is required.This physical problem may be succinctly posed as aradiative transfer equation, the solution of which con-volved with an instrument's response function is theremote spectral measurement.

In this paper, the state-of-the-art on modeling ofradiation transport in vegetation canopies and the appli-cation of such models to the interpretation and analysisof remotely sensed optical data is summarized. Webegin with a statement of the physical problem and itsmathematical description, parameterization and numeri-cal solution (section II). This is followed by a briefcatalogue of the various distortionary effects inherentin remote observations (section III). The extraction ofinformation (namely estimation of vegetation properties)from remote optical measurements is discussed in thelast section. This paper is not intended to be a revieweither of literature or of the various ideas currently invogue. Instead, it is aimed with the broader scope ofitemizing and elaborating three issues in the opticalremote sensing of vegetation-modeling, caveats, andalgorithms.

II. THE PHYSICAL PROBLEM

In this section we begin with a brief statement of thephysical problem encountered in the optical remotesensing of vegetated land surfaces. Forward modelingof the physical problem is developed as boundary valueproblems in radiative transfer. The parameterizationand solution of the resulting equations are discussedalong with results of validation of the models.

Statement of the Physical ProblemThe atmosphere is illuminated at the top by monodirec-tional solar radiation that is absorbed and / or scattereddepending on the composition of the atmosphere andwavelength of the incident beam. The vegetation canopy

0034-4257 / 95 / $9.50SSDI 0034-4257(94)00073-V

**

Page 2: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

1 70 Myneni et al.

receives both monodirectional and diffuse radiation (dueto collisions in the atmosphere), which it reflects backinto the atmosphere according to its bidirectional re-flectance distribution function. The remote measure-ments are typically the angular and spectral distributionsof radiations exiting the atmosphere. This physical prob-lem may be succinctly posed as a radiative transferequation that describes the interaction of photons inthe atmosphere-vegetation-soil medium, the solution ofwhich is the remote spectral measurement. In someinstances, the spectral radiance field immediately abovethe canopy is measured. The corresponding radiativetransfer problem is also outlined in the following sec-tions.

Forward Problem FormulationWe consider the problem of sensing a vegetated landsurface through an atmosphere and are concerned pri-marily with top-of-the-atmosphere (TOA) spectral radi-ance fields. We shall denote this as the remote sensingproblem, as opposed to the vegetation canopy problemwhere both direct sunlight and diffuse skylight are inci-dent and top-of-the-canopy spectral radiance fields areof interest.

Remote Sensing ProblemThe radiance distribution measured by a sensor of uni-form spectral response function aboard a satellite canbe simulated as the solution of the following radiativetransfer problem. For illustrative purposes, the atmo-sphere is assumed to be horizontally homogeneous, offinite optical depth T, illuminated spatially uniformlyon top (T = 0) by monodirectional solar radiation of inten-sity incident along Q ( < 0), and bound at thebottom (r = TA) by a horizontally heterogeneous vegeta-tion canopy. In the absence of polarization, frequencyshifting interactions and emission, which we assumethroughout this presentation, the steady state monochro-matic radiance or intensity distribution function I(TP,2) is given by the boundary value problem

(+ 1)lr,2) = -id (la)I(O,Q) = Io(Q - Qo)4 ji<0, (lb)

1(rA,jitC) =- ± dQ'Rv7(jtQ (2)| vI(TA7jJQŽ2l7t 2n

p' < O. > . (IC)

Here co, is the single scattering albedo, P, is the rota-tionally invariant scattering phase function for photonscattering from the direction ' into and Rv isthe vegetation canopy bidirectional reflectance factor(BRF). The unit vector Q(,(p) has an azimuthal angleV measured anti-clockwise from the positive x-axis thatis directed North, and a polar angle = cos-' u with

respect to the outward normal (opposite to the -axis,which is directed down into the atmosphere); conse-quently, u c (0,1) denotes upward directions of photontravel and vice versa. The operator (Qet) denotes thedirectional derivative along Q in (rp) space, with P - x,y.

From a numerical point of view it is preferableto separate the uncollided from the collided radiationintensity (Myneni et al., 1990). This we do by letting

lr fQ) j= IS(2 Q) + PI(TQ) (2)

and substituting in Eq. (1) to result in the following:

(A) Uncollided Intensity

(Q V, + )l(rjQ) = 0,

P(OQ) = 1o6(Q - o),

(3a)

(3b)p<0,

I(TApSQ) = - dQ 'Rv(p W' " 04u'II"(0,Q)I 2n

P'<O, ji>0,

(3c)

the solution of which is

I(0 r&,) = I(0,Q)exp - i-il, jt<0,

l (rj,2) = lo(rapŽ,Q)exp[- (T|:A-

(4a)

,u > O. (4b)

(B) Collided Intensity

(not, + 1)=(rji,) (OA- dP (Q'OQ)P(rfiY)

+ j dPa(Q')Io(T$4,?) (5a)47 4r

P(OQ) = 0, p < 0, (5b)

IP(rAP,Q) =- | dQRv(#, Q' Q) lv If(TAQ),

p'<0, p>0. (5c)

Problem (B) can be numerically solved by any of thestandard methods in radiative transfer for the collidedradiation intensity P (for example, by the discrete ordi-nates method; cf. Myneni et al., 1990). The atmosphericparameters TA, WOA and PA can be specified for modelatmospheres or from measurements (section II, Atmo-spheric Parameters).

The surface bidirectional reflectance factor Rv thatappears in Eqs. (3c) and (5c) is defined as

Rv(l5,f2' - itI(=nkr"AQ) ta' < 0,Ii''I(Ta,

u > O, (6)

where l(Ta, Q') is the intensity incident on the canopyalong ' and I(TA,P,Q) is the surface radiance along 2.

x exp [ - F"1U_1 I I I-4

Page 3: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 1 71

The latter is obtained from a numerical solution of thecanopy radiative transfer equation.

The classical approach to modeling radiative trans-fer in vegetation has been to ignore all plant organsother than leaves, and treat this leaf canopy as a gaswith nondimensional planar scattering centers, that is,a turbid medium (Ross, 1981). Such an analogy permitsthe use of standard theory, with minor modifications toaccount for the angular orientation of the plane-scatterers.The parameter optical depth depends on the directionof photon travel and the scattering phase functions arenot rotationally invariant, thereby precluding the use ofpolynomial expansion methods for handling the scatter-ing integral (Shultis and Myneni, 1988; Marshak, 1989).

A vegetation canopy may be realistically idealizedas aggregations or clumps of leaves distributed randomlyin free space (vacuum). The intervening free spacesbetween the clumps constitute the voids. The voids areconvolutely shaped and multiply connected three-dimen-sional structures broken along those regions where leavesare present. Consequently, a leaf canopy can be ab-stracted as a binary medium-randomly distributed leafclumps and voids (Myneni et al., 1991). So, the idea ofa continuum host material that is central to radiativetransfer is not provided for in leaf canopies at opticalwavelengths.

If size of the scatterers is considered in a formula-tion of the transport problem, it is necessary to includenot only the number density of scatterers but also theconsequences of introducing finite size gaps (holes orvoids) in the medium. Standard transport descriptionbased on interaction coefficients derived from elemen-tary volumes is not applicable because of the presenceof voids. Moreover, if the far-field assumption is violated,as is the case in a leaf canopy, then the scattererswill cast shadows; hence, information on the spatialdistribution of scatterers is required to evaluate crossshadowing (Myneni and Asrar, 1991).

A need for such a theory has a strong basis in experi-mental observations of exiting radiations measured inopposition to a monodirectional source (Kuusk, 1985).The classical theory assumes an infinite number of un-correlated nondimensional scatterers thereby affording acontinuum description of the configuration space. It isclear that point scatterers cannot cast shadows andthus, the classical theory fails to predict or duplicateexperimental observations of exiting radiations, espe-cially about the opposition direction. The theory ofphoton movement and interactions in media where thescattering centers are finite dimensional-oriented plates,spatially distributed in clumps, with large interveningfree spaces has been developed in a series of recentworks (Myneni et al., 1991; Myneni and Asrar, 1991;Myneni and Ganapol, 1991; Knyazikhin et al., 1992).

A numerical solution of the transport problem in-cluding scatterer size effects has not yet been attempted

because of the difficulty of parameterization and com-plexity of the equation set. Methods based on the ex-change of radiosity between finite dimensional scatter-ers provide an alternative to the study of this difficultproblem (Borel et al., 1991; Goel et al., 1991). Forremote sensing purposes, where the emphasis is onphysically realistic but simple models, heuristic formula-tions generally suffice (Kuusk, 1985; Jupp et al., 1986;among several others). Of these, the analysis of Ver-straete et al. (1990) is especially notable for a precisegeometric formultion of the scatterer size effects. Oneadvantageous strategy is to include scatterer size effectsin first scattering and resort to classical (point scatterer)theory for multiple scatterings (Marshak, 1989), as de-scribed below.

A horizontally heterogeneous vegetation canopy ofphysical depth Zc bounded by a flat and anisotropicallyreflecting soil surface is considered. Specifically, weseek a numerical solution to the following boundaryvalue problem:

[GeV + u(r,)]I( ) dQ'a,(r,2'-2)1(?,2-4,

I(go) = 6(2 - a),v T 1

(7a)

v < 0, (7b)

I(ZcpQ)= dQ'Rs,2' Q)ju'7 2n

x (Zc'PQ)' '< ) u> 0.

(7c)

The above assumes a spatially uniform incidence of unitflux at the top of the canopy (z = 0), and that Rs is thesoil BRF distribution. For convenience we have retainedthe same geometry as earlier [Eq. (1)] but with thez-coordinate replacing the T-coordinate, and -(x,y,z).Here, a is the extinction coefficient and a, is the differ-ential scattering coefficient that depends on the absolutedirections of photon travel ' and Q. The rotationallyinvariant problem is recovered if the slope distributionof scattering centers is random.

Extinction of radiation in a canopy depends on theleaf area density distribution and leaf normal orientation,that is, the slope distribution. It is independent of wave-length. A probabilistic description of the scattering eventrequires information on the leaf-scattering physics inaddition to the above two structural parameters. Simplemodels of the leaf-scattering phase function describingspecular reflection at the leaf surface and diffuse scatter-ing in the leaf interior are available in the literature(Marshak, 1989).

A model for the extinction coefficient that correlatesinteraction rates between incident and once-scatteredphotons is used here to describe the hot-spot effect(Stewart, 1990)

Page 4: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

1 72 Myneni et al.

- exp[ - ED(Q,Q]}, if (.25 < 0,(ar(2Q) if (5> 0, (8)

where E is an empirical parameter related to the ratioof vegetation height to characteristic leaf dimension. Itsvalue was estimated to be between 1 and 8 based onseveral sets of experimental data (Stewart, 1990). Thecorrelation distance D is

D(Q A)= [1 +1 +2(92*1) 0

This particular model for the modified extinction co-efficient has two desirable features, namely, that for( = - 2', a vanishes to result in the hot-spot effect, andfor large scattering angles it approaches the standardcoefficient a. Moreover, it is always positive unlike themodel proposed by Marshak (1989). The inclusion of din the transfer problem [Eq. (7)] is discussed below.

As in the previous case, let

ife,_) 1°(?,2) + t(,Q) + l1(r,Q) + I(rtQ) + I(F2)

(9)

where I is the uncollided intensity, 11 is the once-collided intensity due to downward incident radiation,and I" is the remainder intensity. Inserting Eq. (9) intoEq. (7) results in the following:

(C) Downward Uncollided Intensity

[Q*V + (r,)]Wr,2) = 0, P < 0,

n(OQ) = n -Q- ), U <0,

IY(-rQ= I voexp[- -1 i dsu(o - s'2 oo)Jl

,,<O.

(D) Upward Uncollided Intensity

[2. + &(A,25]U(?,2) = 0,

(Zc,42) =-Rs(5,92 -) iuOIl?(Zc,i,5Q),

v>0, v, <0,

IV(r,Q) = (Zcp2)exp - I 0 ds'T(

,u>O.

(bOa)

(lOb)

11(0,2) = 0, v < 0,

ItRA) = dsfexpL3 dsa(? - si"2,)j

(12b)

v<0, vo<O.

(12c)

(F) Upward Once-Collided Intensity due to I

[29V + &it,2,Q]Nl2,Q) = a 8 -t 2 Q

v>0, v,,<O,

I1(ZC,p,Q) =0, v > 0,

(13a)

(13b)

1 r -dr 1 -Fdlf~r,Q) = I' ds'exp[ 1 Ar dsW( - s'K2Qg,Q,)

x ai,2j - Q)-(pQ2),v>0, v, <O.

(13c)

(G) Remainder Intensity I"

[29V +a(t,2)]Fu(t2)= 4d2'a4A2' 2 0)`Q,25

+ i d4'o t' n )1(r,)

+ d2'a n (a),2n ~~~~(14a)

It(O,Q) = O. A < O, (14b)

I-(Zc,,) =-| d_'Rs(j,2 -2) l/2 I(Zc,Q)2 2)

+-i.2 dQ'R,(P,Q'- ) p'l F(&~,Q'),

v'<0, v>0. (14c)

(10c) In the above, it - (,x,y) is a point on the upper surfaceof the canopy and likewise Ac- (Zc,x,y) on the lower(soil) surface. The variable s denotes distance back alongthe direction such as = - sQ. It is convenient from

v >0, (Ila) a numerical point of view to solve problems (E) and(F) using the discrete ordinates method rather thannumerically evaluating Eqs. (12c) and (13c) because theintensity P' is required to specify the source term [Eq.

(lib) (14a)] and boundary condition [Eq. (14c)] in problem(G). The transfer problems defined by Eqs. (5), (12),

- s'Q,Q,Q4, (13) and (14) can all be solved by the standard discreteI ordinates method (Myneni et al., 1990) or by iterative

methods (Knyazikhin and Marshak, 1991) with some(lie) minor modifications.

(E) Downward Once-Collided Intensity due to I Vegetation Canopy Problem

+ a 2)]1i(?,2) = Ox(A, D 2)2(A20 ) The canopy radiative transfer problem differs from Eq.[2eV + = a,(Y',Q0 - (7) in that both direct sunlight and diffuse (collided in

v < 0, v0< 0, (12a) the atmosphere) skylight are incident on the canopy.

X U,(PQ� - frIVQ ),

Page 5: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 1 73

The atmospheric radiative transfer problem [Eq. (1)]must be solved first in order to specify the boundaryconditions for the canopy problem. In this section, abrief description of the canopy problem is given.

The governing radiative transfer equation and bound-ary conditions are of the form

[2eV + a(,2)]I(tQ) = i dQ'a)(Ir,(Q' ? 92),I2

I(0,fl,2) = I~eXP _ TA ( -.

+ Id(Op,Q), v < 0,

I(Zc,P,92)= ± dQR (f,Q2 Q)Lu'In 2r

x (Zcji,2, v' < 0,

v>0,

(15a)

(15b)

(15c)

where IE is the extra-terrestrial solar radiation incidentat the top of the atmosphere along 2, about wavelengthA. The term Id(O,i,Q) is the atmospheric intensity distri-bution incident on the canopy and is obtained from anumerical solution of the transfer problem denoted byEq. (5), that is, the atmospheric transmittance. Theatmospheric transmittance can be spatially averagedwithout loss of much accuracy. With this simplification,the canopy radiative transfer problem [Eq. (15)] can bereduced to the following problems [cf. Eq. (9)]:

(H) Downward Uncollided Intensitu

[QV + a(, Q)]1?&,Q) = 0, v<0,

l.(OQ) = IEexp - 1]-6(!--Q )

+ Id(O,Q), v < O,IA°r,2) = R(, )

1 r-tix exp - i dsa( _ s' CL -l I I

,u < O. (16c

(I) Upward Uncollided Intensity

[Q*V + (,4)]I(,2) = 0, v > 0, (17a

AI(Zc,,92) = I191 (2)Kl2ilI4(ZcP,2i),

u>0, vi<O, (17b

ItjCr,2) = 1A(ZcA)exp[ - l C ds'&(? -s'2,n,2)1

v>0.

Ir,Q) = ZROF11a), i=0,1,...,N

(J) Downward Once-Collided Intensity due to I

[Q2V + a2)]t(t2) = X

v<0, v,,<O,I(0,Q) = O. u < O,

(18a)

(18b)

Il(r,2) = ds'exp [l- T dsla(r -s'9,Q)]

27t

v<0, v, <O. (18c)

(K) Upward Once-Collided Intensity due to

[2eV + &(t,2Q)]1hit,2) = a (7 (-Q)I?~t,2i),

> v. pi<O, (19a)

Al(ZcPQ) = 0, > 0, (19b)

A (r,2) = J ds'exp[ Mt ds"&(r-s"Qn)

x C,(r,Gi i - Yr'Q)

v>0, vi<O.IVr,2) = ZAi(rF,2),

(19c)

i = 0,1,. .. ,N

(L) Remainder Intensity P'

Same as Problem (G)-Eq. (14)

In the above equations, Id is the spatially averaged(16a) atmospheric transmittance incident along Q with i = 0,

1,2,. . .,M, where 2M is the number of discrete directionsin the unit sphere. The index i = 0 corresponds to theincident solar direction Q. with the assumption that

(16b) ,, $ Q,, where QN is the quadrature set of discreteordinates. A particular advantage of this formulation isthat it circumvents the ambiguity associated with thecoefficient 6, which has a dependence on the photon'sprevious direction of travel [namely Eqs. (17) and (19)].

Parameterization, Solution, and ValidationIn the section entitled "Forward Problem Formulation"we have seen how the forward modeling of the physicalproblem can be posed as boundary value problems inradiative transfer. A numerical solution of these equa-tions requires parameterization of the composition andoptical properties of the media in question. This issueis addressed here together with some examples of modelvalidation.

Atmospheric ParametersAtmospheric optical depth TA, single scattering albedoW0A and phase function PA are parameterized for stan-dard, cloudless, and horizontally homogeneous atmo-

I I

Page 6: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

1 74 Myneni et al.

spheres. Molecular optical depth rA at wavelength A isevaluated from refractive index of air and moleculardensity distribution. Routines from the 5S code can beused to calculate Tt and the Rayleigh phase function Pm(Tanr6 et al., 1990). The profile of the atmosphericaerosol distribution is based on a recommendation bythe International Radiation Commission (IRC) applica-ble to continental areas with an aerosol optical depthof 0.23 at 0.55 vm (Deepak and Gerber, 1983). Theaerosol scattering phase function P' can be modeledusing the Henyey-Greenstein function or other suchanalytical expressions. The asymmetry parameter of theHenyey-Greenstein function g and the aerosol singlescattering albedo coa are tabulated in the IRC reportfor several wavelengths in the solar spectrum. The totalatmospheric optical depth TA, single scattering albedoW9A and the scattering phase function PA are evaluated

assuming external mixing. A modified version of LOW-TRAN-7 is used to select wavelength bands where atmo-spheric absorption is less than 0.1 (S. C. Tsay, personalcommunication). Both line and continuum absorptionby 16 gaseous species at any solar zenith angle can beconsidered in selecting these windows (Table 1).

Canopy ParametersNumerical evaluation of the coefficients a and a, inEqs. (7a) and (15a) requires information on the leaf-areadensity distribution in the stand UL, leaf normal orienta-tion distribution gL and the leaf scattering phase functionYL. A vegetation canopy can be simulated as clumps ofleaves randomly distributed on a reflective soil with aground cover (ge) between 0 and 100 %. A flat horizontalground area (a,) of dimensions Xs x Ys is considered forsimulation purposes. The height of the canopy is as-sumed equal to the height of the clumps (Zs = Zc). If L,is the clump leaf-area index, then by definition, thecanopy leaf area index is given by the product g x LcLeaf-area density distribution inside the clumps canbe modeled using simple analytical functions, namelyquadratic (Myneni et al., 1990). For instance, if leaf area

inside a clump is assumed to be uniformly distributed,then UL = L, I Zc.

The leaf normal orientation can be assumed to beazimuthally symmetric. The distribution along the polarangle can be modeled using the standard distributions:planophile, erectophile, plagiophile, extremophile, anduniform. The discrete Beta distribution is recommendedbecause of its simplicity and versatility (Strebel et al.,1985). With only two parameters this distribution repre-sents most leaf normal orientations encountered in na-ture.

The leaf scattering phase function 2k, can be writtenas the sum of diffuse scattering inside the leaf (YLD)

(the bi-Lambertian model; Ross, 1981) and specularreflection at the leaf surface (YLS) (Vanderbilt and Grant,1985). The leaf spectral reflectance (rL) and transmit-tance (tL) are typically measured with integratingspheres. They can also be simulated using models ofradiation transport in leaves. Jacquemoud and Baret(1990) developed one such model that allows simulationof rL and tL in the 0.4-2.5 vm interval, using only threeparameters-a parameter characterizing leaf mesophyllstructure, chlorophyll concentration, and leaf water con-tent. Table 2 presents values of standard leaf opticalproperties for the wavelength bands of interest.

Soil ParametersBidirectional reflectance factors of the soil, Rs, can bedescribed following Hapke's formulation applied to barerough soil surfaces (Pinty et al., 1989; Jacquemoud et al.,1992). The soil is assumed to be a half-space (semi-infinite) and first-collision intensity is evaluated analyti-cally, with shadows included to model the hot spoteffect. Multiple scattering is simplified to a two-streamproblem. Standard values for the coefficients (b, c, b,c') of the Legendre polynomial expansion for the particlephase function, and the parameter h related to theporosity of the medium, are given in Jacquemoud et al.(1992). The single scattering albedo of the soil particu-

Table 1. Atmospheric Parameters Used in the Base Case

Atmospheric parameters

AerosolIncident Rayleigh Aerosol single

Wavelength energy optical optical scat. Anisotropic(micrometers) (w/ms) depth depth albedo parameter

0.4011-0.5133 214.48 0.200 0.282 0.899 0.6420.5153-0.5333 37.00 0.114 0.242 0.896 0.6380.5353-0.5873 102.15 0.086 0.225 0.892 0.6370.5893-0.6852 163.57 0.051 0.194 0.887 0.6330.6912-0.6972 14.49 0.036 0.175 0.880 0.6310.8280-0.8940 69.62 0.015 0.132 0.842 0.632

Within these wavelength bands atmospheric absorption is less than 10%. A clear and turbidatmospheric conditions were simulated by doubling and halving the total atmospheric opticaldepth in each wavehand.

Page 7: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 1 75

Table 2. Canopy and Soil Parameters Used in the Base Case

Canopy and soil parameters

Single scattering albedoLeaf Leaf of soil particulatesWavelength hemispherical hemispherical

am) reflectance transmittance Sand Clay Peat

0.4011-0.5133 0.058 0.008 0.470 0.217 0.0610.5153-0.5333 0.108 0.061 0.445 0.198 0.0450.5353-0.5873 0.131 0.084 0.536 0.214 0.0540.5893-0.6852 0.083 0.038 0.600 0.267 0.0740.6912-0.6972 0.091 0.051 0.612 0.271 0.0810.8280-0.8940 0.510 0.418 0.710 0.316 0.194

The leaf optical properties are average values of several measured spectra (F.G. Hall, Per-sonal communication). The soil values are taken from Jacquemoud et al. (1992).

lates cos is the only parameter dependent on wavelengthand soil moisture (Fig. 4 in Jacquemoud et al., 1992),although this assumption is not necessarily beyond de-bate. The single scattering albedo of clay, sand, andpeat particulates, integrated over the wavebands of in-terest, are given in Table 2.

Numerical SolutionThe radiative transfer problems [Eqs. (1) and (15)] canbe simplified and approximated under a variety of as-sumptions. The medium may be assumed to be horizon-tally homogeneous to permit a one-dimensional analysis.The transport equation can be azimuthally averaged toresult in a one-angle problem (Shultis and Myneni, 1988)or integrated over all directions to obtain flux transportequations (Suits, 1972). The medium can be assumedto be semi-infinite to allow substantial simplifications(Verstraete et al., 1990). The leaf-normal orientationcan be assumed to be a delta function to facilitate semi-analytical solutions (Ganapol and Myneni, 1992). Multi-ple scattering can be approximated using solutions offlux propagation equations (Nilson and Kuusk, 1989).Similarly, assumptions about the boundary conditions(vacuum and/ or isotropic boundary conditions) facili-tate numerical solution of the transport problem (Dick-inson et al., 1990). Detailed reviews of literature rele-vant to optical remote sensing problems can be foundelsewhere (Goel, 1988; Kaufman, 1989; Myneni andRoss, 1991).

The standard discrete ordinates method with somemodifications can be employed to numerically solve thetransfer problems (B), (E)-(G), (J)-(L), for the radiationintensity (AQ). The discrete ordinates algorithm satis-fies photon conservation and guarantees positivity, ho-mogeneity and stability of the solution (Knyazikin andMarshak, 1991).

The angular variable is discretized using equalweight (EQN) quadrature sets of order N (or other quad-rature schemes), thereby restricting the flight of photonsto discrete quadrature ordinate directions. The spatial

derivatives are approximated by finite difference schemes(or, finite elements). The resulting algebraic system ofequations are solved iteratively on the scattering sourceby the method of sweeping in the phase-space grid.Convergence of this iteration is slow in optically deepcanopies and in situations where the single scatteringalbedo is close to unity. Hence, it is desirable to acceler-ate convergence of the iteration, for which several stan-dard methods exist (Myneni et al., 1990).

The calculation scheme is as follows. The canopyBRF matrix Rv is evaluated first [Problems (C)-(G)]. Theatmospheric transfer problem is then solved using thismatrix [Problems (A)-(B)]. The radiation field incidenton the canopy is now known [Eq. (15b)]. So, the canopytransport problem can be solved [Problems (H)-(L)]. Adetailed energy balance is performed at each level.Interpolation schemes to decouple quadrature ordersbetween the canopy and atmospheric radiative transfercalculations can be implemented for tailoring accuracyand also to evaluate the radiance field in desired angulargrids. Essentially this involves performing one additionalsweep of the phase-space, using the converged sourcedistribution, into the desired angular grid. This entirescheme is wavelength specific. Details on the applica-tion of the discrete ordinates method for numericalsolution of the transport equation can be found in My-neni et al. (1990).

Model Comparison and ResultsA comparison of four different models with varying as-sumptions, but all based on a numerical solution of thetransfer equation, is shown in Figure la. The implemen-tation of our discrete ordinates method was rigorouslybenchmarked and found to be four-digit accurate (My-neni et al., 1988; Ganapol and Myneni, 1992). Themethod was also validated with experimental data ofreflectance spectra collected over maize and soybean(Myneni et al., 1988; Shultis and Myneni, 1988; Stewart,1990; Myneni and Asrar, 1993), prairie grassland (Asraret al., 1989) and forest canopies (Myneni et al., 1992).

Page 8: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

1 76 Myneni et al.

1.0

0.8

0.61

Principal plane

0.41-

80 60 (B) 20 0 20 (F) 60Viewing angle (degrees)

C

4-

00

80

0

-20

-40

80 60 (B) 20 0 20 (F) 60Viewing angle (degrees)

xLt = 1.000

BO 60 (B) 20 0 20 (F) 6viewing angle (degrees)

a,

I--

UCa)

V

V

0 80

10

5

-5 *.~-... ..--10 /- -

-15'80 60 (B) 20 0 20 (F) 6

Viewing angle (degrees)

xLt = 1 0.000

Discrete ordinates

- - Nilson & Kuusk, 1989

...... .Pinty et al., 1990

-.- loquintaFigure la. Comparison of canopy bidirectional reflectance factors in the solar principal plane as predicted by differ-ent models. The canopy leaf area index (xLt) is 1 and 10. The leaf normal orientation is erectophile in polar angleand uniform in azimuth. Leaf single scattering albedo is 0.9 and scattering from soil is assumed Lambertian with ahemispherical reflectance value of 0.3. Henyey-Greenstein phase function with a backscattering asymmetry parameter(g= - 0.4) was used in all the models. The hot spot effect was suppressed in all the models because of different pa-rameterizations. The solar zenith and azimuth angles are 45° and 180°, respectively. The discrete ordinates model isof Shultis and Myneni (1988) and is used here as a reference. The other models are described in Nilson and Kuusk(1989) and Pinty et al. (1990). The model of aquinta (unpublished) combines the first scattering of Verstraete et al.(1990) with the two-stream model of Sellers (1985).

0L_

To

U

0

0

C0

._

a)

C0U

L.

m

I. . .

--- .7-

.Z

/

*1 /., /-'I

80

0t-To0G) 1.0U0U 0.8

V

0' 0.6

0I 0.4

._

.

)

m2 0 80

Page 9: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 1 77

The discrete ordinates model will be used as the bench-mark. The model of Pinty et al. (1990) is notable for itconsiders a canopy of infinite thickness. All of the mod-els except the discrete ordinates use a flux approxima-tion for multiple scattering. Of these, it appears that amodel based on an accurate evaluation of first scatteringand the two-stream model of Sellers (1985) for multiplescattering gives the best approximation to the solutionof the transfer equation (the unpublished model of Ia-quinta). In general, results of comparison of model pre-dictions with empirical data, although satisfactory inmany cases, reveal two major deficiencies in most mod-els; namely, the effects of clumping and size of canopyelements. A realistic treatment of these effects and asimplified three-dimensional model of radiation transferare topics for future studies.

For vegetation, one generally deals with wave-lengths in two spectral regions located on both sides of0.7 um. Indeed, the strong chlorophyll absorption thatcorresponds to low values of the single scattering albedo(less than 0.2) should allow us to retrieve both theoptical and structural properties of the medium becauseof the resulting high anisotropy in the exiting radiancefield. The near infrared region is characterized by highleaf reflectance and transmittance, allowing radiation toeasily reach the ground after multiple scattering events.Moreover, the exiting radiance field is less anisotropicbecause photons are diffused in all directions due tomultiple collisions. This means that the retrieval of

canopy properties will be more difficult. Thus, the ratioof single to total scattering is an indicator of the canopybehaviour with respect to its intrinsic radiative proper-ties (Figure lb). The ratio is strongly nonlinear withrespect to leaf area index and the single scatteringalbedo. It can be seen that the single scattered radianceis less than 50% of the total radiance for the largepart. This clearly indicates that the multiple scatteringcomponent has to be modeled rather accurately forextracting quantitative information from data collectedover this single scattering albedo-leaf area index domain.

CAVEATS IN REMOTE SENSING

The central hypothesis in remote sensing is that radia-tion reflected from a surface carries information aboutthe state of the surface. The appropriation of signalcharacteristics with surface properties is complicatedfor one can never measure in isolation the desiredcause and effect in the system. In the remote sensingof vegetation from satellite-borne instruments, severalcaveats are in order. Several of these are discussed inthis section.

Bidirectional Effects

The angular distribution of top-of-the-canopy (TOC)spectral radiance is shown in Figure 2a for two wave-lengths and solar zenith angles. The canopy leaf areaindex is 3 and uniform leaf normal orientation distribu-

Figure lb. The ratio of single to total scattered radiance obtained from a two-stream model (Sellers, 1985). The il-lumination angle is 45°, the soil albedo is 0.3 and the leaf normal orientation is uniform.

0.

N .

Page 10: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

1 78 Myneni et al.

tion is assumed. The other parameter values are givenin Tables 1 and 2. At both wavelengths, the distributionsare anisotropic with the hot spot in the retro-solar direc-tion. The hot-spot radiance is, in general, the maximumof the distribution. The minimum is found near thenadir on the forward scattering side. The hot-spot effectis more pronounced in the red than near-infrared be-cause of the contrast between sunlit and shaded ele-ments. The degree of anisotropy increases with solarzenith angle. From these results, it can be seen thatdifferences in radiance observations measured at differ-ent solar and view angles are not necessarily attributableto changes in the surface. The differences may be dueto bidirectional effects of the radiance field.

Atmospheric Effects

The corresponding angular distribution of top-of-the-atmosphere (TOA) spectral radiance field is shown inFigure 2b. TOA radiances differ from those above thecanopy both in terms of their magnitude and angulardistribution. The net atmospheric effect, which is thedifference between TOA and TOC values, decreasesalmost linearly with increasing surface reflectance (Kauf-man, 1989). It is positive at shorter wavelengths (namelyred) where atmospheric scattering plays a dominant

Figure 2a. Angular distribution of top of the canopy radi-ance in the solar principal plane at red and near-infraredwavelength. Unit incident irradiance is assumed such that ra-diance x 7r is the corresponding bidirectional reflectance fac-tor. 06 is the solar zenith angle in degrees.

TOP-OF-THE-CANOPY RADIANCE

role and negative at longer wavelengths (near-infrared)where aerosol and gaseous absorption predominate. Forinstance, TOA radiance at the red wavelength is 2 to10 times greater than TOC red radiance depending onthe view and solar zenith angles. The hot-spot effectso distinctive at this wavelength in the TOC radiancedistribution is masked due to scattering in the atmo-sphere. At near-infrared wavelengths, the net atmo-spheric effect is only slightly negative. The TOA radi-ance distribution at this wavelength is similar to thatabove the canopy. When remote sensing vegetationthrough an atmosphere, these spectral atmosphericeffects must be corrected.

Canopy Structural EffectsThe architecture of a leaf canopy, for the purposes ofstudying radiation transport, can be characterized bythe canopy height, leaf size, orientation and densitydistribution. Leaf size distribution and canopy heightdetermine the amplitude and width of the vegetationhot spot (Marshak, 1989). The influence of leaf orienta-tion on TOC spectra is shown in Figure 3, togetherwith the spectrum of a leaf simulated with the PROS-PECT model of Jacquemoud and Baret (1990). If aparameter such as the leaf area index is to be determined

Figure 2b. Angular distribution of top of the atmosphere ra-diance in the solar principal plane at red and near-infraredwavelength. Unit incident irradiance is assumed such that ra-diance x 7t is the corresponding bidirectional reflectance fac-tor. 06 is the solar zenith angle in degrees.

. ..... . .. ..

near-infrared

Oo = 60 ...................

Oo= 15 ,I , I~~~~~~~~~~~~~~~,in red I

0o= 60 - I,

-50.0 0.0 50.0Backscattering Forward Sci

VIEW ZENITH ANGLE (degrees)

TOP-OF-THE-ATMOSPHERE RADIANCE

0.5

a

0

i

A:

0.1 _

0.05 _

0.02m

ittering

Oo= 15, . . ...... o - . . .........,

A~~~~~~~~~~...... .. ................ .

near-infrared ...

;O ..........--- , -"'''

red

-50.0 0.0 50.0Backscattering Forward Scattering

VIEW ZENITH ANGLE (degrees)

0.2

0.1 H

a.4-

E0

2:a

0.05 _

0.02 _

0.01

0.005 [-

0.2 _-

.1 - I nn

Page 11: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 1 79

from remote radiance observations, the structural effectdue to leaf orientation must be kept in mind.

Background or Soil Effects

The radiance field measured above the vegetated landsurface is a composite of both soil and vegetation contri-butions. When vegetation parameters are to be deter-mined from TOC measurements, methods to minimizethe contribution of the soil must be developed (Huete,1988). On the other hand, when surface parameterssuch as the albedo are of interest, it is important toaccurately include soil contribution. This is illustratedin Figure 4, where surface albedo is plotted againstcanopy leaf area index for different canopy parametersand for the cases of a bright and a dark soil. Soil brightnessdetermined the nature of the relationship between sur-face albedo and canopy LAI. When bare soil and densecanopy albedo are comparable in magnitude, as in brightsoil simulation (0.27 and 0.30, respectively), surfacealbedo is a weak function of leaf area index. On theother hand, if the soil is absorptive (bare-soil albedois 0.06), surface albedo increases near-linearly forLAIs < 3, after which it tends to an asymptote. Hence,the surface albedo of a vegetated land surface can varywidely depending upon soil brightness.

Nonlinear Effects of ScatteringThe radiative transfer problem is linear in incident radia-tion, but nonlinear in scattering. The consequences of

non-linearity are illustrated in Figure 5, where thechange in TOC nadir bidirectional reflectance factor(BRF) is plotted against the single scattering albedo(sum of leaf reflectance and transmittance) at two valuesof canopy leaf area index (LAI). In the case of a sparsecanopy (LAI = 1), changes in leaf reflectance (ArL = 0.05& 0.1) at any value of the single scattering albedo arenot detectable in TOC nadir BRFs. On the other handfor a dense canopy (LAI = 5), changes in leaf reflectanceare detectable in TOC BRFs, which depend non-linearlyon the single scattering albedo. These effects are im-portant in the remote sensing of leaf biochemical constit-uents, where changes in constituent concentration trans-late into changes in leaf optical properties.

Effects of Spatial HeterogeneitySpatial distribution of the radiance field and associatedprocesses is affected by spatial heterogeneity of the vege-tated surface. Heterogeneity can be conceptually quanti-fied at a macroscopic scale by ground cover and plantspatial distribution. The structural and optical propertiesof plants in a given area are spatially distributed undernatural conditions. As a first approximation, these prop-erties may be assumed spatially invariant; that is, theplants in a given area are identical. In this case, spatialheterogeneity can be prescribed by ground cover, plantspatial distribution, and plant leaf area index. The impor-tance of spatial heterogeneity arises in the context ofscaling point observations to areal values.

Figure 3. Spectral distribution of nadir vegetation bidirec-tional reflectance factors for canopies with different leaf nor-mal orientations. The leaf area index of the canopies is 3.The solar zenith and azimuth angles are 300 and 0. Thecorresponding leaf hemispherical reflectance spectrum isalso shown.

0.5l

0.4L

U

0.3 4

0.2

0.17-

0.0-

0.5

- 0.4

0.3

0.2

0.1

-0.0

1.0 2.0 3.0

C9

2:

C2:Z

M.

Figure 4. Surface albedo and canopy leaf area index fordark and bright soils, and planophile and erectophile cano-pies. Leaf hemispherical reflectance and transmittance at vis-ible wavelengths are denoted as rL and t, respectively.

0.35

0 .3 0r

005

4 0.20

y 0.15

0.10-

planophile

------- -::7_ - -2(rL+tL)

,;-: bright soil

- r. 0.5(r -L)-

erectophi 7

I .

,'In

,' very dark soil

0.05F-

0.0

0.0 2.0 4.0 6.0 8.0 10.0

CANOPY LEAF AREA INDEX

0.0

II

-

WAVELENGTH (microns)

Page 12: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

180 Myneni et al.

0.5F

0.4

0.3

0.2

0.1

0.0

0.2 0.4 0.6 0.8

LEAF REFLECTANCE + TRANSMITTANCE

Figure 5. Change in vegetation canopy nadir bidirectionalreflectance factor with single scattering albedo (sum of leafhemispherical reflectance, r, and transmittance, tL).

The relationship between the fraction of photosyn-thetically active radiation absorbed by the photosynthe-sizing tissue in a canopy (FAPAR) and the bidirectionalreflectance transform NDVI is an example of scale in-variance (Figure 6). It can be seen that the relationshipbetween FAPAR and TOC NDVI is similar for homoge-neous (ID) and heterogeneous (3D) canopies. There isa near-unique correspondence between TOC NDVI andFAPAR irrespective of the spatial distribution of leafarea. A given value of TOC NDVI might result fromvarious configuations of ground cover and plant leafarea index. In all such cases, FAPAR is nearly unique.This relationship is scale invariant, for pixels of differentspatial scales but of equivalent values of TOC NDVIare likely to have equivalent FAPAR values.

Adjacency Effects

The radiance field measured by a remote sensor maycontain contributions from areas adjacent to the targetof interest because of scattering in the atmosphere intothe field of view. The adjacency effects result in loss ofcontrast when ground areas of varying reflectivities areobserved. This is illustrated in Figure 7, where thesquare wave modulation transfer function (MTF) is plot-ted against spatial frequency. The MTF of an opticalsystem describes the variation in output contrast be-tween the peaks and valleys of a sinusoidal or square-wave pattern as a function of spatial frequency (Dinerand Martonchik, 1984),

1.0 -

< 0.8

C 0.6 -

0cc

00-< 0.4 -4.

Is 0.20.

0.0l

Planophile--iD

---- Planophile--3D.Erectophile--iD

.............. Erectophile--3D

0.0 0.2 0.4 0.6 0.8 1,0

TOC NDVI

Figure 6. The relationship between the fraction of photosyn-thetically active radiation absorbed by vegetation (FAPAR)and top-of-the-canopy normalized difference vegetation in-dex (TOC NDVI) in homogeneous (ID) and heterogeneous(3D) canopies with different leaf normal orientation distribu-tions.

Figure 7. The square-wave normalized modulation transferfunction of strip field ground albedo pattern for three atmo-spheres evaluated from nadir intensities using Eq. 20.

1.00

0.g0

++ --.,, *Turbidity = 04.' ' s.

_ \ ~~~~~~~~~~~~... ..... .... .. . .. . .. .

+' Turbidity = I

Turbidity = 3

0.60 N

CZW

0.40 -Wavelength = 0.55 MicronsSolar Zenith Angle = 40 Degrees

..... Monte Carlo (Pearce, 1977)

0.20* Discrete Ordinates

FREQUENCY (CYCLES/KM)

go

E5

3

I I I

Page 13: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 181

MTF(AX) =

1 [I(AX 2,y,OQ) - I( - AX / 2,y,0,Q)MTFOA I(AX I2,y,0,Q) + I( - AX /2,y,0,Q)j

(20)

where AX is width of the strip, and MTFO is the MTFas AX-oo. The adjacency effects reduce the contrastbetween alternating bright and dark strips with increasein atmospheric optical depth and frequency of groundreflectivity variations.

Nonlinear MixingThe radiance measured by a remote instrument can bewritten as the sum of component radiances because thegoverning equation is linear in radiance. For instance,the total radiant intensity can be written as the sumof unscattered, once-scattered, and multiply-scatteredradiant intensities. With respect to the problem parame-ters, such as incident and observation angles, leaf areaindex (or, optical depth), canopy single scattering albedoand soil reflectance, the equation is nonlinear becauseof the multiple scattering term. The nonlinear effectdue to scattering between leaves and soil is particularlyrelevant in this context. Consider, for instance, a pixelof incomplete ground cover. The radiance of this pixelis not a linear combination of the vegetation and soilradiances because of the radiative interaction betweenthe vegetation and soil due to multiple scattering. Thisis illustrated in Figure 8 where the pixel radiance atnear-infrared is plotted against the radiance at red. Thelinear mixing model assumes that vegetation and soilradiances can be combined linearly with weights propor-tional to ground cover. This assumption is clearly vio-lated under natural conditions because of multiplebounces of photons between leaves and the soil. Thus,care must be excercised when multispectral data areanalyzed to derive information on the scene componentsbased on linear mixing models.

Topographic EffectsIt has long been recognized that topography can signifi-cantly modulate remotely sensed radiances, leading toerrors in retrieved surface reflectances. For an unvege-tated sloping surface, topography affects the estimationof reflectances from TOA radiances in two ways. First,the amount of energy incident on a surface is a functionof topography (via slope, aspect, and shadowing). Sec-ond, the upwelling flux off the surface is also dependenton the position of the slope relative to the sensor.Topographic factors that must be considered includeelevation, slope, aspect, local shadowing by nearby ter-rain, obstruction of diffuse sky radiation by nearby ter-rain, and the configuration of nearby terrain relative tothe slope of determining the amount of energy reflectedtowards the slope. The problem is confounded becausetopographic effects are dependent on the waveband of

0 F4 Full c n p

0.3-

U0. Full Canol

0.2 -

aZ

2: -

Bare Soil

RED REFLECTANCE

Figure 8. Relationship between canopy nadir bidirectionalreflectance factors at red and near-infrared wavelengths.Canopies of varying leaf area indices were simulated bychanging ground cover (0-100%) and plant leaf area index(PLAI = 1 & 5). The nonlinear mixing results were obtainedfrom a solution of the three-dimensional radiative transferequation.

interest and on the resolution of the sensor. If the IFOVincludes many different terrain elements (resolutionsgreater than a hundred meters or so) each element hasits own topographic effect and contribution to the totalradiance. This creates a very difficult inversion problem.A summary of topographic effects and some correctionstrategies can be found in Dozier (1989) and Dubayah(1992).

Existing correction models are probably of limitedapplicability for vegetated surfaces. The elements of acanopy are rarely oriented parallel to a slope; therefore,the effect of the underlying slope is not clear. Research isneeded to couple topographic formulations with canopyradiative transfer models. Such a coupled model couldbe used toward the development of a systematic ap-proach for correcting radiances from vegetated surfacesfor topographic effects.

ALGORITHMS

Algorithms in remote sensing are tools that connectradiance measurements to surface properties of interest.In vegetation remote sensing one can make a distinctionbetween radiation model parameters (leaf area index,leaf and soil optical properties, etc.) and surface stateparameters (spectral albedos and radiation absorptionamounts, photosynthetic and stomatal conductance rates,etc.). The corresponding relations in many cases are

Page 14: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

1 82 Myneni et al.

complicated nonlinear functions (Sellers, 1985) and wemust resort to substantial simplifications in developingalgorithmss if they are to be practical. In this section,we shall discuss some of the methods that are currentlyin use and that may be broadly termed as algorithms.

Spectral Vegetation IndicesSpectral Vegetation Indices (SVI) are synthesized fromspectral reflectance factors using a variety of techniques(Tucker, 1979). A number of SVIs have been proposedin the literature that employ various combinations (sum-ming, differencing, ratioing, etc.) of vegetation Bidirec-tional Reflectance Factors at two or more wavelengths.It is not our objective to review the various indices orperform a comparative analysis here; the discussion islimited to a few indices and their utility in inferring thedesired surface properties.

The most common indices are the NormalizedDifference Vegetation Index and the Simple Ratio (SR).The component canopy reflectance factors invariablycontain contribution from the soil or background litter.A Soil-Adjusted Vegetation Index (SAVI) was proposedby Huete (1988), which minimizes soil brightness influ-ences on SVIs involving red and near-infrared wave-lengths. In addition to minimizing the effect of back-ground, top-of-the-atmosphere spectral radiance valuesmust be corrected for atmospheric effects to recover thevegetation signal. One such index, the AtmosphericallyResistant Vegetation Index (ARVI), was recently pro-posed by Kaufman and Tanr6 (1992) and incorporatesa self-correction process for the atmospheric effect atred wavelength by utilizing the radiance difference be-tween blue and red wavelengths. The Global Environ-ment Monitoring Index (GEMI) was recently developedto specifically correct for atmospheric effects in AVHRRdata by employing a nonlinear combination of red andnear-infrared reflectances (Pinty and Verstraete, 1992a).GEMI exhibits a high atmospheric transmissivity, insen-sitivity to soil reflectance, with the exception of verybright soils, and is empirically representative of vegeta-tion properties in a manner similar to the other indices.

There is substantial empirical, and in some casestheoretical, evidence that these indices are related toseveral vegetation parameters such as ground cover, leafarea index, radiation absorption, canopy photosynthesis,canopy conductance, etc. (Hall et al., 1992). Althoughthese findings are highly encouraging, it must be empha-sized that the information content of an index dependsonly on the constituent radiances (or reflectances) andthat an empirical relationship does not necessarily implya causal relationship. Vegetation indices are a conve-nient way of summarizing the remote observations andare useful in mapping and event detection, and in someinstances of estimating surface parameters, an exampleof which is discussed below.

From theoretical analysis it can be shown that thenormalized contrast in canopy reflectance between redand near-infrared wavelengths (NDVI) is, within somebounds, responsive to the leaf area in a canopy. Simi-larly, the amount of photosynthetically active radiationabsorbed by the photosynthesizing tissue (FAPAR) isrelated to the green leaf area of the canopy. Thus, wemay expect a causal relationship between NDVI andFAPAR (Asrar et al., 1984). The nature of this relation-ship and how it varies with respect to changes in canopy,soil, and atmospheric parameters was investigated utiliz-ing the radiative transfer method outlined in the sectionentitled "Statement of The Physical Problem" (Myneniand Williams, 1993). The relationship between FAPARand NDVI was found to be independent of pixel hetero-geneity, parameterized with ground cover, clump leafarea index, and variations in leaf orientation and opticalproperties. On the other hand, it was sensitive to back-ground, atmospheric, and bidirectional effects. Atmo-spheric and bidirectional effects may be ignored ifthe analysis is limited to near-nadir top-of-the-canopyNDVI. Further, if the soils are moderately reflective,background effects may also be ignored. A linear modelfits the resulting relationship between FAPAR and nadirTOC NDVI very significantly (r2 = 0.943, N = 280). Theslope of this relationship is 0.8465 with an intercept of- 0.1083 (cf. Myneni and Williams, 1994). This linear

model or algorithm for FAPAR is valid for: (a) solarzenith angle less than 60°, (b) view zenith angles aboutthe nadir, (c) soils or backgrounds of moderate bright-ness (NDVI about 0.12), (d) atmospheric optical depthsless than 0.65 at 550 nm.

Ecosystem productivity models can be driven withFAPAR estimates derived from remote radiance obser-vations. Spectral measurements from satellites must befirst corrected for pixel location and converted to radi-ances using proper geometric restitution and calibrationprocedures. The TOC radiance field can then be de-duced from the TOA distribution using an atmosphericcorrection algorithm. This has been done for 13 AVHRRimages of Belgium by Veroustratete et al. (1993). TheTOC NDVIs thus derived were converted to FAPARvalues using the linear model described above. Such aFAPAR image of Belgium is shown in Figure 9. Afine resolution FAPAR time series restituted from thecoarser empirical series through Fourier analysis wasthen utilized to drive a model of net ecosystem carbonexchange in a deciduous forest (Veroustraete et al., 1993).

Model InversionA successful model inversion allows the retrieval ofseveral independent model parameters. An inversionmay further allow the calculation of surface state vari-ables (spectral albedos, FAPAR, etc.) from a limited setof empirical reflectance values depending on the stabil-

Page 15: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 183

FAPAR - 8890iBfp..img13/07/90

UUU

iU

.

11

I_

a -t~

Ndl,

E

-9. coo2726

-E C00023

-7.0000906

0.00002 . Coo0

7220. 250

4353

7130

113420.3125

.m75011925

. 3706 24

0.50003555

0 .0250o179

0.7500

6555

2 .0CC

4.10

s 66

1 . 35

21 .46

1s 19

9.2C0

5.47

3 .27

. 0cC

1, 0 . 0

Figure 9. Map of fraction of photosynthetically active radiation absorbed by vegetation derived from atmospherically cor-rected AVHRR NDVI data.

ity and uniqueness of the problem. Although parametersof biological or climatological importance may be mostdesirable, practical limitation imposed by the numberand distribution of reflectance samples, spectral regions,signal anisotrophy, and model sensitivity may dictatethe optimal parameter set.

In order to infer any vegetation parameter withgood accuracy, model inversion needs to be mathemati-cally well conditioned; it implies that the partial deriva-tives of the bidirectional reflectance field with respectto a particular parameter has to be nonzero and shouldtake values as high as possible. Goel and his colleaguespioneered the study of model inversions as detailed ina series of papers beginning in 1983. After developingan appropriate problem definition, merit function andinverstion strategy (Goel and Strebel, 1983), they in-verted several one-dimensional models-most notablythe Suits (1972), SAIL (Verhoef, 1984) and Cupid (Nor-man, 1979) models-and quantified parameter relation-ships and model sensitivity. Strahler and his colleaguesdeveloped and inverted a reflectance model that simu-lates a heterogeneous vegetation canopy as an assort-

ment of geometrical objects (Li and Strahler, 1985).Antyufeev and Marshak (1990) developed a radiativetransfer model and inverted it using Monte Carlo tech-niques. They retrieved three optical and four geometri-cal canopy parameters from reflectance data at twowavelengths. The hybrid three-dimensional model de-veloped by Welles and Norman (1991) combines thegeometrical objects and radiative transfer approachesto simulate the reflectance of heterogeneous vegetationcanopies. This model has been inverted by Goel some-what less successfully (Goel, 1992, personal communica-tion). Recently, simple analytical models have been de-veloped and inverted specifically for application toclimate models (Dickinson et al., 1990; Pinty et al.,1990). General discussion on model inversion and somecentral ideas are concisely summarized in Privette etal. (1994).

The general inversion problem may be stated asfollows: given a set of empirical directional reflectancemeasurements, determine the set of independent modelparameters such that the computed directional re-flectances best fit the empirical directional reflectances.

Page 16: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

184 Myneni et al.

The fit of the empirical data is determined by the meritfunction (Goel and Strebel, 1983), E2, defined as

(21)n

Q:2= (rj _ rn)2j=l

where rj is the directional reflectance for a given scanand solar angle geometry, rn,,, is the geometrically analo-gous model estimate, and n is the number of reflectancesamples. A penalty function may be used to limit theindependent parameter space to physically possible val-ues. For instance, the sum of leaf reflectance and trans-mittance can be constrained, based on physical argu-ments, to be 1. The ability to correctly determinetarget parameters through model inversion thereforedepends on the dataset r 1 , the likeness of the model tophysical reality and ability of the optimization algorithmchosen to minimize Eq. (21) over the parameter space.

There are several methods for minimizing the meritfunction and the choice of a particular method dependson the mathematical properties of the function to beminimized. Three commonly used minimization rou-tines are: the downhill simplex method used by Privetteet al. (1993) (subroutine AMOEBA from Press et al.,1986), the conjugate direction set method used byKuusk (1991) (subroutine POWELL from Press et al.,1986), and a quasi-Newton method used by Pinty et al.(1990) (subroutine E04JAF from Numerical AlgorithmsGroup). These three routines do not require evaluationof function derivatives for minimization. This is particu-larly appealing for complex, nonlinear formulations, suchas those used in canopy reflectance modelling.

In practice, measured data are contaminated withnoise from various sources. In the context of modelinversion, the question of the impact of noise on theaccuracy of the parameters retrieved must be addressed.These effects can be analyzed using synthetic data sets,that is, model-generated data, where a Gaussian noiseof zero mean and known standard deviation is added tothe data (Pinty et al., 1990).

As mentioned above, model inversions are per-formed using Eq. (21) with an optimization routineprovided by standard libraries [E04JAF from the Numer-ical Algorithm Group for solving the nonlinear systemwith a quasi-Newton scheme, and RAN 1, GASDEV fromNumerical Recipes (Press et al., 1986) for the generationof random noise]. A number of synthetic noisy data setscan be created depending on the conditions used toselect the random noise series. The retrieved values foreach of the model parameters can be averaged, whichleads to an evaluation of the behavior of the inversionprocedure with respect to noisy data sets. A critical issuewhen performing such an analysis is the randomness ofthe series of generated numbers, which in turn dependson the seed values used in the procedure. To illustratethis aspect, we considered two different cases by chang-ing the seed values applied to the Numerical Recipes

20.08

8 20.07

l 20.06"A

20.05

20.04

20.03

s 20.02

20.01

0.5 1 1.5 2 2.5 3 3.5 4 4.5

Standard deviation of noise x10-3

Figure Oa. Mean values of the single scattering albedo as afunction of the standard deviation of the noise in the inputdata. For a given level of noise, each value corresponds toan average of 100 inversions performed by changing theseed according to a regular progression of negative integernumbers (Case 1). The same noise generator is used withchanging values in the standard deviation of the noise.

8

3 20.J

0 20.05'a0

;~ 205

Standard deviation of noise x10-

Figure 10b. Same as Figure 10a except that the seeds arerandomly distributed both for a given level of noise andstandard deviation of the noise (Case 2).

algorithm-the seeding series given by a regular pro-gression of negative integer numbers (Case 1) and by aseries of a priori random integer values obtained froma different external routine (Case 2). Figures lOa andlOb show the results of these two methods for a sensitivevegetation parameter, the single scattering albedo. Themean quantities reported are averages over 100 individ-ual values, each of these corresponding to the retrievalmade using the same level of noise but with a differentseed as indicated above. The results obtained in Case1 indicate a systematic overestimation of this modelparameter with increasing noise level and, the smooth-

Page 17: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 185

x10-3 E041AF -Smooth initialization of noise

1 2

Standard deviation of noise

4

xt 13

Figure 1c. Distribution of the standard deviations corre-sponding to the mean single scattering albedos obtainedwhen performed 100 inversions, given as a function of thestandard deviation of the noise in the input data (Case 1).

Table 3. True and Mean Retrieved Parameter Valueswith Standard Deviations for Soybean Data of 27 Aug.1980.

True Retrieved

Parameter p , a

LAI (both channels) 2.9 4.1 2.5LAI (red only) 5.4 2.4LAI (NIR only) 2.8 1.9Leaf refl. (NIR) 0.454 0.495 0.067Leaf refl. (red) 0.073 0.089 0.006Leaf trans. (NIR) 0.518 0.376 0.055Leaf trans. (red) 0.064 0.070 0.028Soil refl. (NIR) 0.232 0.418 0.127Soil refl. (red) 0.143 0.133 0.045,u (Beta LAD) 1.61 1.54 0.502v (Beta LAD) 2.18 2.52 1.03

Unless otherwise noted, retrieved values were averaged over all 12datasets and both channels. True LAD parameter values are from Goeland Strebel (1983).

6

2

01 2 3 4 5

Standard deviation of noise x10-3

Figure 10d. Same as Figure 10c; (Case 2).

ness of the plotted curves is clearly noticeable. By contrast,Case 2 leads to much more variable and nonmonotonicfeatures regarding the relationship between the noiselevel and the average retrieved single scattering albedo.For the same two cases, Figures lOc and lOd show thestandard deviation corresponding to each of the tenretrievals made at a given noise level, plotted as afunction of this noise level. Both figures depict a quasi-linear relationship between the standard deviation inthe input noise and the standard deviation in the re-trieved single scattering albedo values. In Case 2, thedistribution of the standard deviation values at a givenlevel of input noise appears to be more noisy than inCase 1.

Inversion of empirical data with a one-dimensionalradiative transfer model (Shultis and Myneni, 1988) is

illustrated with the soybean data collected by Ransonet al. on 27 August 1988 (Privette et al., 1993). Datafrom two spectral bands, 0.6-0.7 pm (red) and 0.8-1.1,um (NIR), were used in the inversion. The downhillsimplex method was used for minimizing the meritfunction. The mean and standard deviations of the re-trieved parameters are shown in Table 3. The meanretrieved LAI is high, although the measured value iswithin one standard deviation. Closer inspection revealsthat the red LAI estimate is high but the NIR estimateis excellent (0.1 absolute error). The coefficients of theBeta distribution are close to the true values and suggesta predominantly erectophile canopy. Estimates of leafreflectance were reasonable though slightly high. Al-though the red reflectance value differed by just 0.016absolute, the measured value was outside the one stan-dard deviation confidence interval. The NIR value differedby 0.041 but was within the confidence interval. Leaftransmittance was overestimated by 0.006 in the redand underestimated by 0.142 in the NIR. The red soilreflectance estimate was excellent (5% relative error).The NIR estimate was high (80% relative error).

At the present time, inversion of physical models ismore of an art than exact science. Several topics ininversion studies need to be investigated before anoperational scheme can be proposed. Some of theseinclude (1) alternate forms of the merit function andtheir mathematical properties, (2) optimum number ofparameters for retrieval, (3) optimal sampling schemes,(4) the importance of noise in the measurements, (5)success in parameter retrieval at various wavelengths,(6) starting points, (7) nesting of models of varyingaccuracy in an inversion scheme and (8) definition of asuccessful inversion. Furthermore, the competing ad-vantages of computational efficiency and physical realityshould dictate some optimum level of model sophistica-

.0

IWB1

Page 18: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

186 Myneni et al.

tion. This threshold presently has not been determinedand undoubtedly will change with technical improve-ments.

Expert SystemConventional algorithms such as those based on spectralindices for inferring surface hemispherical reflectance,percentage of ground cover, biomass, leaf area index,photosynthetic capacity, etc. are static and limited inscope. These algorithms cannot deal with the variabilityinherent in remote sensing data (e.g., variable off-nadirviewing capabilities, varying solar zenith angles, varioussensor wavelengths, cloud cover problems, missing data,atmospheric effects, etc.). Information extraction sys-tems, at a minimum, must be designed with these factorsin mind. In addition, it is desirable to obtain an estimateof the accuracy of inference. Moreover, systems mustbe flexible enough to handle decisions at the expert'slevel.

In the recent years an expert system called VEGwas developed with the above in mind (Kimes et al.,1991, 1992). The current version of VEG is a powerfulsystem for inferring vegetation characteristics using na-dir and/or directional reflectance data as input. VEGis also a powerful tool for aiding a researcher in devel-oping new and more accurate techniques for inferringvegetation characteristics as well as testing these tech-niques. Finally, VEG has options to process files ofremotely sensed data collected over large regions of theearth.

Part of VEG is designed to have any array of extrac-tion techniques for inferring vegetation characteristicsusing nadir and / or directional reflectance data as input.Currently, the system has many techniques for inferringspectral hemispherical reflectance, percentage of groundcover and view angle extension. It is designed to beeasily expanded to handle other inferences, such astotal hemispherical reflectance, leaf area index, biomass,photosynthetic capacity, etc. The system intelligentlyand efficiently integrates traditional spectral data withdiverse knowledge bases available in the literature, fromfield data sets of directional scattering behavior, andfrom human experts. VEG accepts any combination ofnadir and / or off-nadir spectral data of an unknowntarget as input, determines the best strategy(s) for infer-ring the desired vegetation characteristic, applies thestrategy(s) to the target data, and provides a rigorousestimate of the accuracy of the inference. Several re-search applications using VEG have been made (Kimesand Deering, 1992; Kimes and Holben, 1992; Kimes etal., 1993a, 1993b).

Neural Networks and Genetic AlgorithmsRecently, some new approaches to the extraction ofscene parameters from optical remote sensing data have

been developed by combining physically based modelsand measurements with neural computing techniques.Ishimaru et al. (1990) outlined the basic approach ofemploying a backpropagation neural network to invertatmospheric particle size distribution parameters fromoptical data. Smith (1993) first trained a backpropaga-tion neural network to invert a simple multiple scatter-ing model to estimate leaf area index from reflectanceat three wavelengths and then subsequently applied thetrained network to satellite observations. He reportedestimated errors of less than 30% and indicated thatthe method appeared to be much less sensitive to initialguesses for the parameters than other inversion tech-niques. A potential drawback to the technique is thedifficulty of interpreting the network weights, althoughSmith (1993) discussed the use of a genetic algorithmto map the decision boundaries of the inversion.

CONCLUDING REMARKS

Models of varying degree of sophistication describingmost of the known physical mechanisms are availablein the literature. However, research is required in thearea of modeling tree architecture, especially for conif-erous species, and quantifying spatial heterogeneity atthe landscape level. Further simplification of most mod-els is required if they are to be utilized in the analysisof satellite data. The information content of remotelysensed data may be conceptually ordinated along thetemporal, spatial, spectral and directional (and, polariza-tion) dimensions. Algorithms must be devised to exploitthe full information content in the signal. The state-of-the-art is far from satisfactory. In spite of obvious limita-tions, spectral vegetation indices are still preferable inthe analysis of large spatial scale data sets. The promiseof remote sensing, however, lies in those methods thatutilize physical models and advances in computer sci-ence and technology.

This work was made possible with the financial support ofNASA grant NAS5-30442 to RBM and DLW The sectionsBidirectional Effects and Neural Networks and Genetic Algo-rithms are contributions of Drs. R. Dubayah and J. A. Smith,respectively, and we gratefully acknowledge their participation.JLP is a DOE Global Change Graduate Fellow. Most of thismanuscript was prepared at the Blaise Pascal University inFrance; JLP and RBM acknowledge the hospitality and supportof the Laboratoire de Mgttorologie Physique in Clermont-Ferrand.

REFERENCES

Antyufeev, V. S., and Marshak, A. L. (1990), Inversion ofa Monte Carlo model for estimating vegetation canopyparameters, Remote Sens. Environ. 23:201-209.

Asrar, G., Fuchs, M., Kanemasu, E. T., and Hatfield, J. H.(1984), Estimating absorbed photosynthetic radiation and

Page 19: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

Optical Remote Sensing of Vegetation 187

leaf area index from spectral reflectance in wheat, Agron.J. 76:300-306.

Asrar, G., Myneni, R. B., Li, Y., and Kanemasu, E. T. (1989),Measuring and modelling spectral characteristics of a tall-grass prairie, Remote Sens. Environ. 27:143-155.

Borel, C. C., and Gerstl, S. A. W., and Powers, B. J. (1991),The radiosity method in optical remote sensing of struc-tured 3-D surfaces, Remote Sens. Environ. 36:13-44.

Deepak, A., and Gerber, H. E. (Eds.) (1983), Report ofWMO (CAS)/Radiation (Williamsburg, Virginia, USA, 28-30 March, 1983), World Meteorological Organisation, Ge-neva, Report No. WCP-55.

Dickinson, R. E., Pinty, B., and Verstraete, M. M. (1990),Relating surface albedos in GCM to remotely sensed data,Agric. For. Meteorol. 52:109-132.

Diner, D. J., and Martonchik, J. V. (1984), Atmospheric trans-fer of radiation above an inhomogeneous Non-Lambertianreflecting ground. II. Computational considerations andresults,J. Quant. Spectroscp. Radiat. Transfer, 32:279-304.

Dozier, J. (1989), Spectral signature of alpine snow cover fromthe Landsat Thematic Mapper, Remote Sens. Environ. 28:9-22.

Dubayah, R. (1992), Estimating net solar radiation using Land-sat Thematic Mapper and digital elevation data, Water Re-sources Res. 28:2469-2484.

Ganapol, B. D., and Myneni, R. B. (1992), The FN method forthe one-angle radiation transfer equation applied to plantcanopies. Remote Sens. Environ. 39:213-231.

Goel, N. S. (1988), Models of vegetation canopy reflectance andtheir use in estimation of biophysical parameters fromreflectance data, Remote Sens. Rev. 4:1-222.

Goel, N. S., Rozenhal, I., and Thompson, R. L. (1991), A com-puter graphics based model for scattering from objects ofarbitrary shapes in the optical region, Remote Sens. Environ.36:73-104.

Goel, N. S., and Strebel, D. E. (1983). Inversion of vegetationcanopy reflectance models for estimating agronomic vari-ables. I: Problem definition and initial results using theSuits' model, Remote Sens. Environ. 13:487-507.

Hall, F. G., Huemmrich, K. F., Goetz, S. J., Sellers, P. J., andNickeson, J. E. (1992), Satellite remote sensing of surfaceenergy balance: Success, failures, and unresolved issues inFIFE, J. Geophys. Res. 97:19,061-19,089.

Huete, A. R. (1988), A Soil-adjusted vegetation index (SAVI),Remote Sens. Environ. 25:295-309.

Ishimaru, A., Marks, R. J., Tsang, L., Lam, C. M., and Park,D. C. (1990), Particle size distribution determination usingoptical sensing and neural networks, Opt. Lett. 25:1221-1223.

Jacquemoud, S., and Baret, F. (1990), PROSPECT: A modelof leaf optical properties spectra, Remote Sens. Environ. 34:75-91.

Jacquemoud, S., Baret, F., and Hanocq, J. F. (1992), Modellingspectral and bidirectional soil reflectance, Remote Sens.Environ. 41:123-132.

Jupp, D. L. B., Walker, J., and Penridge, L. K. (1986), Interpre-tation of vegetation structure in Landsat MSS imagery: acase study in disturbed semi-arid eucalypt woodlands. Part2. Model-based analysis. J. Environ. Manage. 23:35-57.

Kaufman, Y. J. (1989), The atmospheric effect on remote sensingand its corrections, in Theory and Applications of Optical

Remote Sensing (G. Asrar, Ed.), J. Wiley, New York, pp.336-428.

Kaufman, Y. J., and Tanr6, D. (1992), Atmospherically resistantvegetation index (ARVI) for EOS-MODIS, IEEE Trans.Geosci. Remote Sens. 30:261-270.

Kimes, D.S., and Deering, D. W. (1992), Remote sensing ofsurface hemispherical reflectance (albedo) using pointablemulti-spectral imaging spectrometers, Remote Sens. Envi-ron. 39:85-94.

Kimes, D. S., and Holben, B. N. (1992), Extracting spectralalbedo from NOAA 9 AVHRR multiple view data using anatmospheric correction procedure and. an expert system,IEEE Trans. Geosci. Remote Sens. 13:275-289.

Kimes, D. S., Harrison, P. R., and Ratcliffe, P. A. (1991), AKnowledge-based expert system for inferring vegetationcharacteristics, Int. J Remote Sens. 12:1987-2020.

Kimes, D. S., Harrison, P. R., and Harrison, P. A. (1992),Learning class descriptions from a data base of spectralreflectance with multiple view angles, IEEE Geosci. RemoteSens. 30:315-325.

Kimes, D. S., Irons, J. R., and Levine, E. R. (1993a), Learningclass descriptions from a data base of spectral reflectanceof soil samples, Remote Sens. Environ. 43:161-169.

Kimes, D. S., Kerber, A. G., and Sellers, P. A. (1993b), Spatialerrors in creating hemispherical reflectance (albedo) mapsfrom directional reflectance data, Remote Sens. Environ.43:1-10.

Knyazikin, Y., and Marshak, A. L. (1991), Fundamental equa-tions of radiative transfer in leaf canopies and iterativemethods of their solution, In Photon-Vegetation Interac-tions: Applications in Optical Remote Sensing and PlantEcology (R. B. Myneni and J. Ross, Ed.), Springer-Verlag,Heidelberg, Germany, pp. 9-44.

Knyazikhin, Y. V., Marshak, A. L., and Myneni, R. B. (1992),Interaction of photons in a canopy of finite dimensionalleaves, Remote Sens. Environ. 39:61-74.

Kuusk, A. (1985), The hot spot effect of a uniform vegetativecover, Sov. J Remote Sens. 3:645-658.

Kuusk, A. (1991), The determination of vegetation canopyparameters from optical measurements, Remote Sens. Eni-viron. 37:207-218.

Li, X., and Strahler, A. H. (1985), Geometrical-optical model-ling of a conifer forest canopy, IEEE Trans. Geosci. RemoteSens. GE-23:705-721.

Marshak, A. L. (1989), The effect of the hot spot on the transportequation in plant canopies, J. Quant. Spectroscop. RadiatTransfer, 42:615-630.

Myneni, R. B., and Asrar, C. (1991), Photon interaction crosssections for aggregations of finite dimensional leaves, Re-mote Sens. Environ. 37:219-224.

Myneni, R. B., and Asrar, G. (1993), Radiative transfer inthree dimensional atmosphere vegetation media, J. QuantSpectroscp Radiat Transfer, 49:585-598.

Myneni, R. B., and Ganapol, B. D. (1991), A simplified formula-tion of photon transport in leaf canopies with finite dimen-sional scatterers, J. Quant Spectroscp. Radiat. Transfer 46:135-140.

Myneni, R. B., and Ross, J. (eds.) (1991), Photon-VegetationInteractions: Applications in Optical Remote Sensing andPlant Ecology, Springer-Verlag, Heidelberg, Germany.

Myneni, R. B., and Williams, D. L. (1994), On the relationship

Page 20: Optical Remote Sensing of Vegetation: Modeling, Caveats ... · Optical Remote Sensing of Vegetation 1 71 The latter is obtained from a numerical solution of the canopy radiative transfer

188 Myneni et al.

between FAPAR and NDVI, Remote Sens. Environ. 49:200-211.

Myneni, R. B., Asrar, G., and Gerstl, S. A. W. (1990), Radiativetransfer in three dimensional leaf canopies, Trans. TheoryStat. Phys. 19:205-250.

Myneni, R. B., Asrar, G., and Hall, F. G. (1992), A threedimensional radiative transfer model for optical remotesensing of vegetated land surfaces, Remote Sens. Environ.41:85-103.

Myneni, R. B., Gutschick, V. P., Asrar, G., and Kanemasu,E. T. (1988), Photon transport in vegetation canopies withanisotropic scattering: Parts II and IV, Agric. For. Meteorol.42:17-40 & 42:101-120.

Myneni, R. B., Marshak, A. L., and Knyazikhin, Yu. (1991),Transport theory for leaf canopies with finite dimensionalscattering centers, J. Quant. Spectroscp. Radiat Transfer,46:259-280.

Nilson, T., and Kuusk, A. (1989), A reflectance model for thehomogeneous plant canopy and its inversion, Remote Sens.Environ. 27:157-167.

Norman, J. M. (1979), Modelling of complete crop canopy, InModification of the Aerial Environment of Plants (B. G.Barfield and J. F. Gerber, Eds.), ASE Monograph No. 2,St. Joseph, Michigan, pp. 249-277.

Pinty, B., and Verstraete, M. M. (1992a), GEMI: A non-linearindex to monitor global vegetation from satellites, Vegetatio101:15-20.

Pinty, B., and Verstraete, M. M. (1992b), On the design andvalidation of surface bidirectional reflectance and albedomodels, Remote Sens. Environ. 41:155-167.

Pinty, B., Verstraete, M. M., and Dickinson, R. E. (1989), Aphysical model for predicting bidirectional reflectance overbare soil, Remote Sens. Environ. 27:273-288.

Pinty, B., Verstraete, M. M., and Dickinson, R. E. (1990), Aphysical model for the bidirectional reflectance of vegeta-tion canopies-Part 2: Inversion and validation,J. Geophys.Res. 95:11767-11775.

Press, W. H., Flannery, B. P., Teukolsky, S. A., and Vetterling,W. T. (1986), Numerical Recipes, Cambridge Univ. Press,New York, pp. 274-312.

Privette, J. L., Myneni, R. B., Tucker, C. J., and Emery,W. J. (1994), Invertibility of a 1-D discrete ordinates canopyreflectance model, Remote Sens. Environ. 48:89-105.

Ranson, K. J., and Biehl, L. L. (1983), Corn Canopy ReflectanceModelling Data Set, LARS Tech. Rep., Purdue Univ., W.Lafayette, Indiana 47907, USA.

Ross, J. (1981), The Radiation Regime and Architecture of PlantStands, Dr. W. Junk Publ., Den Haag, The Netherlands.

Sellers, P. (1985), Canopy reflectance, photosynthesis andtranspiration, Int. J. Remote Sens. 8:1335-1372.

Shultis, J. K., and Myneni, R. B. (1988), Radiative transfer invegetation canopies with anisotropic scattering, J. Quant.Spectroscp. Radiat. Transfer 39:115-129.

Smith, J. A. (1993), LAI inversion using a backpropagationneural network trained with a multiple scattering model,IEEE Trans Geosci. Remote Sens. GE-31: (In Press).

Stewart, R. (1990), Modelling Radiant Energy Transfer in Vege-tation Canopies, M. S. Thesis, Kansas State University, Man-hattan, Kansas 66506.

Strebel, D. E., Goel, N. S., and Ranson, K. J. (1985), Two-dimensional leaf orientation distributions, IEEE Trans. Geosc.Remote Sens. GE-23:640-647.

Suits, G. W. (1972), The calculation of the directional re-flectance of a vegetative canopy, Remote Sens. Environ. 2:117-125.

Tanr6, D., Deroo, C., Duhaut, P., Herman, M., Morcrette,J. J., Perbos, J., and Deschamps, P. Y. (1990), Descriptionof a computer code to simulate the satellite signal in thesolar spectrum: the 5S code, Int. J. Remote Sens. 11:659-668.

Tucker, C. J. (1979), Red and photographic infrared linearcombinations for monitoring vegetation, Remote Sens. Envi-ron. 8:127-150.

Vanderbilt, V. C., and Grant, L. (1985), Plant canopy specularreflectance model, IEEE Trans. Geosc. Remote Sens. GE-23:722-730.

Verhoef, W. (1984), Light scattering by leaf layers with appli-cation to canopy reflectance, Remote Sens. Environ. 16:125-141.

Veroustraete, F., Patyn, J., Van Rensbergen, J., Kretzschmar,J., and Myneni, R. B. (1993), Coupling of Remote SensingData with a Simple Ecosystem Model to Estimate Photosyn-thetic Assimilation, Report ENE.RA9307, Flemish Institutefor Technological Research, Mol, Belgium.

Verstraete, M. M., Pinty, B., and Dickinson, R. E. (1990), Aphysical model of the bidirectional reflectance of vegetationcanopies. 1. Theory, J. Geophys. Res. 95:11755-11765.

Welles, J. M., and Norman, J. M. (1991), Photon Transportin Discontinuous Canopies: A Weighted Random Ap-proach, In Photon-Vegetation Interactions: Applications inOptical Remote Sensing and Plant Ecology (R. B. Myneniand J. Ross, Eds.), Springer-Verlag, Heidelberg, Germany,pp. 389-414.