. u. .. I 2a.“. 2 _. .L ELL . a . L . .. . $3312 3.3:. 31:: .333. Law... , . L .4 . . . L .19 . 3.2.3.‘ . , 5.95:1: .. L. rain 5.53.21 LU. aganuahw 37....) a thug-34.. 1:. , . .7 v.r>¢.4..::.:b_ , uyz ”Mmfdé when . ”1 m4? 1.. [021.1 ”Mm . .i 1. 537151.)... 3 - imrfl. .v v i .9.un l «a . 2 ‘ .LLL 41531;... LLLE: 3L ”LL. ._ A bk 7.... This is to certify that the dissertation entitled ESTIMATION OF TROPICAL FOREST BIOPHYSICAL ATTRIBUTES WITH SYNERGISTIC USE OF OPTICAL AND MICROWAVE REMOTE SENSING TECHNIQUES presented by CUIZHEN WANG has been accepted towards fulfillment of the requirements for the PhD. degree in Geography o‘- I / MabTProfessor’s Signature 5/;“/c»§/ I Date MSU is an Affirmative Action/Equal Opportunity Institution LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 4 AL.“ - IcI 6/01 c:/ClFIC/DateDue.p65-p.15 ESTIIVIATION OF TROPICAL FOREST BIOPHYSICAL ATTRIBUTES WITH SYNERGISTIC USE OF OPTICAL AND MICROWAVE REMOTE SENSING TECHNIQUES By Cuizhen Wang A DISSERTATION Submitted to Michigan State University in partial fulfillment of requirements for the degree of DOCTOR OF PHILOSOPHY Depaitment of Geography 2004 ABSTRACT ESTIMATION OF TROPICAL FOREST BIOPHYSICAL ATTRIBUTES WITH SYNERGISTIC USE OF OPTICAL AND MICROWAVE REMOTE SENSING TECHNIQUES By Cuizhen Wang Accurate estimates of tropical forest biophysical attributes provide quantitative information in the assessment of human disturbances and the carbon sequestration in global climate change studies. The research of my dissertation is thus to develop new algorithms to estimate tropical forest biophysical parameters (forest fractional cover, leaf area index, structures, and aboveground biomass) using synergy of optical and radar remote sensing imagery and radiative transfer models. The case study is in Mae Chaem Watershed, ChiangMai, Thailand. Ground data were collected during the field trips sponsored by research projects. A linear unmixing model in the vegetation index (MSA V1) domain was built to estimate forest fractional cover with Landsat ETM+ image. The forest fractional cover map was validated using both ground measurements and high-resolution IKONOS images. The estimated fractional cover correlated with the ground-measured fractional cover (R2=O.76) at 32 study sites and correlated with the IKONOS—calculated fractional cover (R2=0.70) at 400 randomly selected locations. The leaf area index was estimated using a modified Gaussian regression model with forest fractional cover results. The model was examined with a x2 goodness-of-fit test. The correlation coefficient between the modeled and ground-measured leaf area index values is 0.90. A microwave/optical synergistic radiative transfer mode] was built to simulate the radar scattering from the forest components. The leaf scattering and its attenuation to the woody components (branches, trunks) were quantified with the leaf area index derived from optical remote sensing data. The forest structural parameters, such as tree height and stand density, were estimated through model inversion with JERS-l SAR and VNIR data. The total root—mean—square error (RMSE) of tree height estimation was 4.6 meter and that of stand density estimation was 300 trees/ha. The stand density estimation did not work in tropical evergreen forests because of its saturation at around 500 trees/ha. In accordance with ground measurements, tree height is negatively correlated with stand density in the study area. The model inversion becomes questionable at mountainous areas with high relief and steep slopes. The aboveground forest biomass is also calculated with allometric equations and the modeled forest structural parameters. The total RMSE is 88 ton/ha. The methods developed in this study could be applied to estimate forest biophysical attributes at regional or global scales. With optical remote sensing imagery, only forest fractional cover and leaf area index could be estimated. When both optical and SAR data are acquired, the forest structural parameters and aboveground biomass can be estimated. These results could provide quantitative information in full carbon accounting in global climate change studies. To my husband Lanwu Zhao and my baby girl Jessie with all my love And to my family in china for their infinite support and encouragement iv ACKNOWLEDGMENTS First of all, I would like to thank Dr. J iaguo Qi, my faculty advisor, for his constant help and encouragements during my graduate studies. Without his academic guidance, it is impossible for me to complete this dissertation and move forward to my academic career. I would also like to thank my committee members, Dr. David Skole, Dr. David Lusch, Dr. Richard Kobe, and Dr. William Salas for their support and help even since the dissertation proposal was emerged. I am also grateful to all the faculty and staff members in Department of Geography, Michigan State University, who help me to possess the knowledge as a geographer as quick as I can. Special thanks to Dr. Randall Schaetzl and Ms. Sharon Ruggles who gave me the strongest support when I was struggling in US as a foreign graduate student. Also, I would like to thank Dr. Mark Cochrine, Walter Chomentowski, Jay Samek, Eraldo Matricardi, Oscar Castaneda and Cameron Williams in the Center of Global Change and Earth Observations (CGCEO) for image processing, technical support and helpful comments on my research. Many thanks go to Ms. Diane Cox and Deana Haner, very wonderful ladies for their assistance in my study and research. I would also express my appreciation to Narumon (Nok) Wiangwang, Pari (Perry) Varnakovida, Chetphong Butthep in CGCEO, and Dr. Charlie Navanujraha, Ms. Siam Lawavirojwong and Woody in Mahidol University, Bangkok, Thailand. Their warrnhearted help and assistance in field works in Thailand will never be forgotten. Finally, my deepest thank to my husband Lanwu Zhao who gave me endless support and love during my graduate study. Also to my little sweetheart Jessie who was born when I began to write my dissertation. Her arrival motivated me to be a successful mother and scientist. Thanks to my father, sisters, and brother in China for their love and many years’ support. This dissertation is also a very special memorial to my loved mother. I am sure she will receive this message in heaven. vi TABLE OF CONTENTS Chapter 1 Introduction ...................................................................................................................... 1.1 Literature Review ............................................................................................... 1.2 Research Objectives ........................................................................................... 1 .3 References .......................................................................................................... Chapter 2 Study Area and Field Measurements ............................................................................ 2.1 Study Area ........................................................................................................ 2.2 Field Measurements .......................................................................................... 2.2.1 First field trip (August 10—18, 2001) ......................................................... 2.2.2 Second field trip (January 20-27, 2002) ................................................... 2.2.3 Field data processing ................................................................................. 2.2.3.1 Forest fractional cover .......................................................................... 2.2.3.2 Stand density ......................................................................................... 2.2.3.3 Aboveground biomass .......................................................................... 2.2.4 Ground data analysis ................................................................................. . 2.3 References ......................................................................................................... 1 Chapter 3 Estimation of Tropical Forest Fractional Cover with Landsat ETM+ and IKONOS Imagery ............................................................................................................................ . 3. 1 Introduction ....................................................................................................... . 3.2 Remotely Sensed Data ...................................................................................... . 3.2.1 Landsat ETM+ image ............................................................................... . 3.2.2 High—resolution IKONOS Images ............................................................. I 3.3 Methods ............................................................................................................. I 3.3.1 A linear unmixing model .......................................................................... I 3.3.2 Optimal vegetation index (V1) .................................................................. I 3.4 Canopy Fractional Cover Analysis ................................................................... I 3.5 Validation .......................................................................................................... I 3.5.1 Validation with ground measurements ..................................................... ; 3.5.2 Validation with l-m IKONOS data .......................................................... ; 3.5.3 Seasonal adjustment of fiactional cover ................................................... . 3.6 Conclusion and Discussion ............................................................................... I 3.7 References ......................................................................................................... I Chapter 4 Estimation of Leaf Area Index with Fractional Cover Data in Tropical Forests ..... ‘ 4. 1 Introduction ....................................................................................................... ‘ 4.2 Experimental Design and Field Measurements ................................................ I 4.2.1 LAI-2000 and fisheye photographs .......................................................... I 4.2.2 LA] ground measurements ........................................................................ I 4.2.3 LAI ~ forest fractional cover relationship ................................................. I 4.3 Model Development and Results ...................................................................... I 4.3.1 A modified Gaussian regression model .................................................... I 4.3.2 LA] estimation ........................................................................................... I vii 4.4 Validation .......................................................................................................... E 4.5 Conclusion and Discussion ............................................................................... E 4.6 Reference .......................................................................................................... S Chapter 5 A Microwave/Optical Synergistic Canopy Scattering Model and its Inversion to Estimate Forest Structure ............................................................................................ 1t 5. 1 Introduction ..................................................................................................... 1 ( 5 .2 Model Development ........................................................................................ 1 ( 5.2.1 Modified Karam-IEM model .................................................................. 1( 5.2.1.1 Soil surface scattering ......................................................................... 1( 5.2.1.2 Leaf scattering ..................................................................................... 1( 5.2.1.3 Branch scattering ................................................................................ 1( 5.2.1.4 Trtmk scattering .................................................................................. 1( 5.2.1.5 Leaf-soil interaction ............................................................................ 1( 5.2.1.6 Branch-soil interaction ........................................................................ 1( 5.2.1.7 Trunk-soil interaction .......................................................................... 1( 5.2.2 Linkage to optical remotely sensed variables ......................................... 11 5.2.3 PDF functions of forest components ....................................................... 1] 5.3 Model Simulation and Validation............................. ...................................... 11 5.3.1 Remotely sensed data .............................................................................. 11 5 .3 .2 Model simulation .................................................................................... l 1 5.3.2.1 Contribution of leaves ......................................................................... 11 5.3.2.2 Contribution of branches ..................................................................... 1] 5.3.2.3 Contribution of trunks ......................................................................... 11 5.3.3 Model validation ..................................................................................... 12 5.4 Model Inversion and Forest Parameters Estimation ....................................... 12 5.4.1 Model inversion ...................................................................................... 12 5.4.2 Forest structural parameters by model inversion .................................... 12 5.4.3 Uncertainty analysis ................................................................................ 12 5.5 Conclusions and Discussion ........................................................................... 12 5.6 References ....................................................................................................... 13 Chapter 6 Aboveground Woody Biomass Estimation with Microwave and Optical Remotely Sensed Data .................................................................................................................... 14 6.1 Introduction ..................................................................................................... 14 6.2 Biomass Estimation with a Simple Regression Method ................................. 14 6.3 Biomass Estimation with Compensation of leaf Attenuation ......................... 15 6.4 Biomass Estimation with Synergistic Model and Allometric Equations ........ 15 6.5 Conclusions and Discussions .......................................................................... 15 6.6 References ....................................................................................................... 16 Chapter 7 Conclusions and Future Envisions .............................................................................. 17 7.1 Concluding Remarks ....................................................................................... 17 7.2 Challenges ....................................................................................................... 17 7.3 Potential Applications for Other Studies ........................................................ 17 Viiii Table 2-1 Table 2-2 Table 4-1 Table 5-1 Table 5-2 Table 5-3 Table 5-4 Table 5—5 Table 6—1 LIST OF TABLES Average values of ground measurements in different forest types. ................ 27 Average values of ground measurements in different forest types when the outlier was deleted. ......................................................................................... 27 Study sites in northern Michigan. ................................................................... 95 System parameters of J ERS—l SAR and NVIR sensors ................................ 132 Input parameters for model simulation. ........................................................ 132 Input parameters for model validation (in addition to ground measurements). ......................................................................................... 133 A set of forest structural parameters for each forest type in the model. ....... 133 Average and standard deviation of modeled tree height, stand density, and the error of model inversion for each forest types. ............................................. 134 Coefficients of 0'0 ~biomass logarithmic curve fitting and their statistic tests. ......................................................................................... 164 ix LIST OF FIGURES Figure 2-1 The study area (Mae Chaem Watershed). ....................................................... 2 Figure 2—2 The DEM data (with hill shade effect) in the study area. ............................... 2 Figure 2-3 Vertical distribution of forest types in the study area. .................................... 2 Figure 2-4 Forest type map in the study area. .................................................................. 2 Figure 2-5 Study sites and ground control points (GCPs) in the study area ..................... 3 Figure 2-6 fisheye picture and its transformation in angular coordinates. ....................... 3 Figure 2-7 Correlation scatterplot matrix of ground—measured data. ............................... 3 Figure 2-8 Normalized average ground measurements in different forest types .............. 3 Figure 2-9 Modified normalized average ground measurements in different forest types (the outlier of dry evergreen site was removed). ............................................ 3 Figure 3-1 ETM+(band4+3+2, 02/02/2000) (a) and IKONS images (band4+3+2, 02/27/2000 and 10/09/2002) (b) in Mae Chaem Watershed. The dots in (b) represent the study sites in two field trips ....................................................... 7 Figure 3-2 DEM data with hillshade effect (a) and topographic correction with Rahman’. BRDF model (b). ............................................................................................ 7 Figure 3-3 Dynamic ranges of the six vegetation indices calculated with simulated spectral data in SAIL model. .......................................................................... 7 Figure 3-4 MSAVI image in Mae Chaem Watershed. ..................................................... 7 Figure 3-5 Forest fractional cover map in Mae Chaem Watershed. ................................. 7 Figure 3—6 Comparison of ground-measured and ETM+ estimated fc: scatterplot (a) and clustered column (b). ...................................................................................... 7 Figure 3-8 Scatter plots of ground-measured and the H0861,)“" [l+®2—2®cos(7r—§)]3/2 1+G p(65’6v’¢):p ) (3'3) where p0 and k are two empirical surface parameters. O is the function parameter controlling the relative amount of forward (0 S O S 1) and backward scattering (-1 S O S 0). The phase angle 5 and geometric factor G can be calculated by cos; = cos 6, cos I9, + sin 65 sin 6,, cosrp; G = \/tan2 65 + tan2 6, — 2tan6$ tang, cosrp. Similarly, the surface reflectance p'(6_; ,6; ,go') in local geometry can be calculated. The correction coefficient p(¢95 ,6v,(p) / p'(6ls' ,6,',,,(p') was then applied to the ETM+ surface reflectance image. 40 The topographically corrected image was much smoother even in areas with large topographic variation. Figure 3-2b showed a subset of the ETM+ image (Band4+3+2) near the peak of Mount Doi Inthanon before and after BRDF correction. It is obvious that before BRDF correction, the topographic effect was dominant creating dark shadows in the image. After correction however, the image was totally flattened and the shadows were greatly reduced. After atmospheric, geometric, and topographic correction, the ETM+ image was ready for forest fractional cover estimation. 3.2.2 High-resolution IKON OS Images For the purposes of spatial comparison and seasonal adjustment, ten IKONOS images were acquired in the study area (Figure 3-1b), covering a total area of 673 km2 in Mae Chaem Watershed. The study sites visited during the two field trips were also covered by these IKONOS images. One IKONOS scene was acquired on February 27, 2000, 25 days later than the ETM+ image acquisition, during the dry season. The remaining IKONOS scenes were acquired on October 9, 2002, during the wet season. All of these IKONOS images were geometrically corrected to the ETM+ image. No atmospheric and topographic correction was done on these IKONOS images because only classification and Visual interpretation were needed for the process. The IKONOS images have four spectral bands at 4-m resolution and one panchromatic band at l—m resolution. For each IKONOS image, the l-m panchromatic band was 41 merged with the 4-m spectral bands to produce a pan-sharpened multi—spectral image at l-m resolution. Since the size of tropical trees is much larger than 1 meter, the pan- sharpened IKONOS image reveals enough details for forest fractional cover calculation. The pan-sharpened IKONOS images were then processed to calculate fc at a spatial unit of 30x30 m, comparable with the ETM+ estimated fc. The IKONOS calculated fc values were assumed as “ground truth” to validate the ETM+ estimation in continuous spatial areas . Since most of the IKONOS imagery was acquired in the wet season, the ETM+ estimated fc from the dry season could be adjusted to the wet season conditions when the vegetation was flourishing. After adjustment, the seasonal variation of fc in deciduous forests was minimized and, therefore, the forest fractional cover map provided a better description of forest integrity in the study area. 3.3 Methods 3.3.1 A linear unmixing model For a surface with multiple components, the macroscopic linear mixing model used to estimate the spectral response can be expressed as: [D] = [RIC] (3-4) where [D] is the data matrix, [R] is the response matrix, and [C] is the eigenvector matrix consisting of the relative contributions of different components. It can also be written as: d... = 2w... (3-5) .i =1 42 where di’k is the measured response of pixel k in wavelength 1', n is the total number of independent reflecting components in this pixel, ’1‘.)- is the response of component j in wavelength 1', and CM is the relative contribution of component j in pixel k. Since the ETM+ image in this study was acquired in the dry season, the open areas in the forests were bare or covered by senescent grasses. Consequently, each pixel of the ETM+ image can be assumed to consist of two components: green tree canopy and open area. The spectral response of open areas is characterized by bare soil and does not change much all over the study area. The senescent grass has a similar spectral response to bare soil and, therefore, is neglected. The reflected spectral values of these two components are independent of each other. Let R be the total spectral response of the pixel at certain spectral band, R the response of tree canopy, and R the response of open area, canopy open Eq.3-5 becomes: R = Rama” fc + ROW (1 — fc) + 8 (3-6) where fc is the forest fractional cover in the area corresponding to one pixel. An error term 8 is introduced to account for some insignificant remaining components within the pixel. In tropical forests during the dry season, the two components (green tree canopy and open area) are dominant and the error term, a, is ignored. Eq.3-6 suggests that at any wavelength, the total reflectance of a mixed pixel is the linear combination of the reflectance from its vegetation and open area, weighted by their percentage cover fc and (l-fc). With Eq.3-6, when R and R are known in an canopy open 43 ETM+ image, the fc value can be easily calculated. However, the reflectance of surface targets changes greatly at different wavelengths. For example, the reflectance of green vegetation is high in NIR, but low in red wavelengths. Even at certain wavelengths, the values of R and R are highly influenced by the vegetation wetness, structure, soil canopy OpCII moisture, and texture (J asinski, 1990). Therefore, it is difficult to choose an optimal wavelength for R and R in Eq.3-6. canopy open A vegetation index (V1) is a mathematical combination of multiple spectral bands, which can suppress the external effects mentioned above (Gutman et al. 1998; Qi et al. 2000). Replacing the spectral response R with V], Eq.3-5 becomes: VI = VI canopy fc + VIope" (1 — fc) (3-7) where V1 is the vegetation index of green tree canopies in one pixel, and V1 is the canopy open vegetation index of open area in that pixel. It should be noted, however, that although both vegetation index and spectral reflectance are descriptions of vegetation properties, they are not linearly related. Eq.3-7 is not directly deducted from Eq.3-6. Instead, it is an approximation by choosing the optimal VI that is most linearly related to the vegetation amount in the study area. 3.3.2 Optimal vegetation index (V1) Several vegetation indices (VIs) have been developed over the past decades. The most commonly applied VIs include the Normalized Difference Vegetation Index (ND V1) (Tucker et a1. 1979); the Soil Adjusted Vegetation Index (SA V1) (Huete et al. 1988); the 44 Modified Soil Adjusted Vegetation Index (MSA V1) (Qi et al. 1994); and some vegetation indices for global applications such as the Enhanced Vegetation Index (E VI) (Huete et al. 1999); the Global Environmental Monitoring Index (GEM!) (Pinty and Verstraete 1992); and MERIS Global Vegetation Index (MG V1) (Gobron et al. 1999). Among these vegetation indices, ND VI is the earliest and the most widely applied vegetation index in remote sensing applications. Most of other VIs are modifications ND V1 in an attempt to depress the influence of atmosphere and surface conditions and to improve its accuracy in extracting vegetation information from remotely sensed data. To select the optimal VI in the linear unmixing model, all these VIs (ND V1, SA VI, MSA V], E VI, GEM], and MG VI) were calculated using simulated surface reflectance in the Scattering by Arbitrarily Inclined Leaves (SAIL) model (Verhoef 1984). The reflectance was simulated as a function of leaf area index (LAI), an indicator of green vegetation abundance. A forest of sparse to moderate density was modeled with LA] values from O to 4.0 at an interval of 0.2, which is similar to most of the forests in the study area. The conditions of dense evergreen forests were not considered because both VIs and forest fc values became saturated when LA] was very high. The same sun-target—sensor geometry as the Landsat satellite was used in the reflectance simulations. As shown in Figure 3-3, the values of all six VIs increase with leaf area index. The E V], SA VI, and GEM] quickly reach their saturation zone when LAI approaches 3.0. Although ND VI has a higher value and increases almost linearly at lower LAI, it saturates at a similar LAI threshold as the other three vegetation indices. MG VI has highest value and 45 does not saturate until LAI = 4.0. But, it increases rapidly when LA] is low, then quickly slows down when LAI is higher than 2.0. The MG VI-LAI curve is polynomial. MSA VI has a more linear relationship with LAI. It saturates only when vegetation is very dense (LAI higher than 4.0). Since we are interested in an index that is most suitable for tropical forests, MSA VI was chosen as the optimal vegetation index in this study because of its sensitivity at higher forest densities. The MSA VI reduces the effect of the soil background by using a soil adjustment function that is determined from local soil reflectance. Unlike the empirical factors in other VIs, the adjustment factor in the MSA VI changes as a function of canopy density and the slope of the soil line (Qi et al. 1994): MSA VI = pm —p,,, (1+L0) (3-8) ION/R +pred +1‘0 where L0 is a soil adjustment function. It is expressed as a function of the soil line slope and the reflectance properties on the ground: L0 : [(pNIR _ pred ) X Slope +1+ ION/R “I” prod I2 _ 8'0 x Slope X (ION/R - pred) (3'9) In our study area, the slope of the soil line is 1.2, calculated with the surface reflectance values of soil surfaces in the atmospherically corrected ETM+ data. In the MSA VI image (Figure 3-4), the tropical evergreen forests have high values (light gray to white). The clear-cut areas, bare soil, and fallowed agricultural fields have low values (dark gray or black), while the dry dipterocarps and mixed deciduous forests are in between. 46 From Eq.3-7,fc can be expressed in the form of the MSA VI: MSA VI — MSA V1, ,, c = p (3-10) MSA VI —MSA VI canopy open Here MSA VI and MSA VI are the endmembers in the linear unmixing model. canopy open MSA V1 is the vegetation index of full-cover green forests with fc = 1.0. MSA VI canopy open is the vegetation index of open area with fc = 0. The grasses and shrubs in open areas are senescent in the dry season and, therefore, MSA V1 is close to the MSA VI value of bare open soil. The two endmembers MSA VI and MSA VI in Eq.3-10 can be identified in the canopy open ETM+ image. A sand mining spot with a smooth light gray tone in the middle left of the image was chosen to represent the open area in the study area. It had been observed from the field trips that the top of Mount Don Inthanon was covered by the tropical moist evergreen forests with very high fractional cover values. A small area with a smooth bright red tone in the middle of the ETM+ image (band 4+3+2) was chosen to represent full-cover green canopy. The mean of the MSA VI values in each area was calculated, and the two endmembers in the linear unmixing model were thus determined: MSA VI = canopy 0.71, and MSA VI = 0.16. These values were used in Eq.3-10 to compute the forest open fractional cover in the study area. 47 3.4 Canopy Fractional Cover Analysis Figure 3-5 is the fc map estimated from the ETM+ image in the study area, binned into 10 groups with an interval of 10%. All pixels with fc value less than 0 or higher than 100% (reaching saturation) were truncated into 0 or 100%, respectively. Non-forest surfaces, such as villages, water bodies, fallowed agriculture fields, and clear-cuts, have a fractional cover lower than 20%. Forests in the watershed, with fractional covers ranging from 20% to 100%, are assigned a gradual color scheme changing from light grayish green to dark green, indicating the increasing forest density. The forest fractional cover map in Figure 3-5 shows a similar pattern with the land cover and forest type map (Figure 1—4). Along with the altitude, the forest types change in a sequence of dry dipterocarps, mixed deciduous, dry evergreen and moist evergreen. In accordance, the fractional cover changes from low values in dry dipterocarps to saturation (100%) in moist evergreen forests. Even at each forest type, the fractional cover changes greatly. The fractional cover for each forest type is also not smoothly distributed. Rather, it is scattered because of various disturbances. Since the ETM+ image was acquired during the dry season, dry dipterocarps, especially at lower elevations, were mostly senescent or leaf-off. The vegetation index values are thus much lower than other forests and, therefore, the fractional cover is lower. Some dry dipterocarp forests even have fractional cover lower than 20%. Aside from the seasonal effect, dry dipterocarp forests suffer from intense human disturbance of burning for agriculture or cutting for firewood (personal 48 communication with local foresters). Selective logging for valuable trees such as teak also used to be a common human activity in the past decades. As a result, most of the dry dipterocarp forests are young second-growth forests with a fractional cover ranges from 10% to 40%. The fc values of the forests are lower in areas close to villages and higher in the mountains. The mixed deciduous forests are found at higher elevations and are less affected by seasonal variation and human activities. Clear-cuts for small area agricultural fields and natural fires are thus the common types of deforestation. The fractional cover of mixed deciduous forests ranges from 40% to 80%. Evergreen forests occupy the highest elevations in the watershed. Human disturbance is much lower than for other forests. Natural fire is thus the major disturbance in these forests. Also, as the result of government policy decisions, some large areas, sometimes the whole slopes, were cleared long time ago and evergreen forests regenerated slowly. The fractional cover of the evergreen forests ranges from 70% to 100%. The fractional cover values of most moist evergreen forests on the top of the mountains are saturated. 3.5 Validation The forest fractional cover estimated with the ETM+ image was validated with both ground measurements and hi gh-resolution IKONOS images. The IKONOS images acquired in the wet season can also be processed to evaluate the seasonal change of the forest fractional cover distribution in the study area. 49 3.5.1 Validation with ground measurements The ground-measuredfc values were assumed to be correct (i.e. ground truth). At each site, ten fisheye pictures were taken along a 150m transect (300-m in second field trip) which was assumed to represent the area of 150x150m (or 300x300m). The fc values calculated with these pictures were averaged to represent the fc in an area of 150x150 m (300x300 m in second field trip). The fc estimated from ETM+ image at each site was averaged using a 5x5 window (150x150 m) to match the area of the ground measurements and to reduce the autocorrelations between pixels. Figure 3-6 compares the ground-measured and ETM+ estimated fc values. To be seasonally matched, only measurements from the dry season were compared. Figure 3-6a shows that the ETM+ estimated fc is correlated with ground measurements (R2 = 0.757). The regression line deviates from the 1:1 line, indicating that the ETM+ estimated fc is different from ground-measuredfc. In sparse forests Where the fractional cover is low, the ETM+ estimation is lower than the ground measurements. Contrarily, the ETM+ estimation is higher than ground measurement in dense forests. The ETM+ estimation matches well with ground measurements in forests with moderately high density, in which the fc is around 70% - 85%. In very dense forests, the ETM+fc estimation saturates while the ground measurements are less than 95%. The difference is also shown in Figure 3-6b. Both the ETM+ estimated and ground- measured fc values increases along the elevation gradient, in accordance with the forest 50 types changing in a sequence of Dry dipterocarps, Mixed deciduous, Pine transition, Dry evergreen, and Moist evergreen. For Dry dipterocarps, the ETM+ estimated fc is much lower than ground measurement. For Mixed deciduous and Pine transition forests, the ETM+ estimated fc is lower than the ground measurements, but the difference is much less than that of Dry dipterocarps. For evergreen forests, the ETM+ estimated fc is a little higher than the ground measurements. The ETM+ estimated fc is saturated in the moist evergreen forests (reaching 100%). The difference between the ETM+ estimated and ground-measuredfc comes from the different processing mechanisms. The ETM+ estimated fc in each pixel is calculated with a linear unmixing model in which the vegetation index is a combined contribution of green canopy and open area. Only green leaves in forests are considered in the model. Therefore, the ETM+ estimated fc is actually the “green” fractional cover. The ground— measured fc, however, is calculated from a binary classification of a hemispherical photograph taken on the ground. All elements except open sky, including green leaves, senescent leaves, stems, branches, and trunks, contribute to the value. In this meaning, the ground estimated fc is the total cover of forest projected area. The measurements shown in Figure 3-6 were made in the dry season when deciduous forests are senescent or leaf-off. Therefore, in dry dipterocarps, the ETM+ estimated fc is low due to a lack of green leaves. The ground-measuredfc, however, is much higher because of the significant contribution from the woody components (Figure 3-6b). In the mixed deciduous and pine transition forests, some species were partially leaf-off and 51 others remained evergreen. Consequently, both the ETM+ estimated and ground- measured fc values are higher. The ETM+ estimated fc is still lower than the ground measurements, but the difference is much smaller. Moreover, the vegetation index of pine trees is generally lower than that of broadleaf trees due to the form of the needle leaves. This also makes the ETM+ estimated fc in the pine transition zones lower than the ground measurements. Both the dry evergreen and moist evergreen forests are much denser than the other forest types of the area. The leaves remain green in. the dry season and, therefore, both the ETM+ estimated and the ground-measuredfc values are high (>90%). In dense evergreen forests, the in-tree gaps become dominant when calculating fc with circular hemispherical photographs using the GLA software. However, these small-size gaps are unrecognizable in the ETM+ image with 30-m resolution. As a result, the ETM+ estimated fc is higher than the ground measurements. In very dense, moist evergreen forests, the ETM+ estimated fc reaches saturation while the ground measurement is less than 95%. Figure 3-6 explains the differences between ETM+ estimated and ground—measuredfc values in each forest type. For fractional cover distribution in a large area with all forest types, Figure 3-6a shows that the ETM+ estimated fc values are highly related to the ground truth. This relationship indicates that the forest fractional cover distribution in the study area can be estimated with ETM+ image with a reasonably high accuracy, especially given the leaf-off issue. This validation, however, was based on ground measurements in only 32 isolated sites. To make a spatial validation in the study area, the 52 high resolution IKONOS images were assumed as ground truth to compare with the ETM+ estimated fc distribution. 3.5.2 Validation with l-m IKONOS data The pan-sharpened IKONOS images have four spectral bands at l-m resolution in which the openings between the trees are visible. In this study, the fc derived from the IKONOS data was assumed to be correct (i.e. ground truth) wherever intensive ground measurements are not available. Each IKONOS image was processed in six steps to calculate fc in a unit area of 30x30 1n: Step 1: Unsupervised classification Based on the gray values in the four spectral bands, the pixels in the IKONOS image were grouped into 50 clusters using the ISODATA unsupervised classification technique. A maximum iteration of 50 in 95% convergence level was applied in the classification. The signature file was thus created for next step. Step 2: Supervised classification With the experience of the field trips, the 50 signatures in the signature file were merged into several signatures based on the land cover types and seasonal variation. With the new signature file, a maximum likelihood supervised classification was made of the IKONOS image and a land cover image was generated. There were 5 classes in the dry season: forest, shaded forest, bare soil, shaded bare soil, and water body. For the wet season, there were 8 classes: forest, shaded/dark forest, bright agriculture/open area, fallowed agriculture/open area, bare soil, shaded bare soil, water body, and cloud/shadow 53 (if any). Figure 3-7 shows the feature spaces of each cluster between band 3 (red) and 4 (NIR). The bright forest and dark forest are combined into one cluster because they are the major concern in this classification procedure. All classes are very separable. Step 3: Majority filtering In the land cover image created in step 2, there were often some isolated classes embedded in large classes. For example, there were agriculture fields with one or two pixels in forests. These pixels were not real classes and should be removed. Using a 3x3 moving window and a majority decision rule, the pixels in the small classes were merged into the larger surrounding classes. Stg; 4: Forest/non-forest class The land cover image was recoded into a 0/1 binary image, assigning forest and shaded (or dark) forest as l and all others as 0. The resulting forest/non-forest image maintained a pixel size of 1 meter. Steg 5: 30x30 matrix convolution The forest/non-forest image was convolved with a 30x30 moving window to match the pixel size of the ETM+ image. The resultant pixel value was a float point value between 0 and 1, which was equal to the forest fractional cover in a 30x30m area on the ground. Step 6: Forest fractional cover Applying the nearest neighbor technique, the image from step 5 was resampled to 30-m pixel size. The result was the IKONOS calculated fractional cover distribution with a unit area of 30x30 m. 54 The IKONOS estimated fc distribution in the study area is an alternative to “ground measurements” to spatially validate the ETM+ estimation. It should be noted that, the IKONOS fc values were calculated with unsupervised and supervised classification techniques and were very different from both hemispherical photographs for ground fc calculation and the linear unmixing model for ETM+fc estimation. The ground-measured and the IKONOS and ETM+ estimated fc values were compared in Figure 3-8. To avoid the seasonal effect, only fc data in the dry season were compared. Only 7 study sites were covered by the IKONOS image in the dry season. All forest types were included: dry dipterocarps (1), mixed deciduous (2), pine transition (1), and evergreen (3). Among these 7 sites, only 6 sites were measured during the second field trip. Therefore, in Figure 3-8a, there are 7 points of ETM+fc, 7 points of IKONOSfc, and 6 points of ground-measuredfc. When compared with ground fc values from fisheye photos, the IKONOS fc was underestimated in sparse forests and overestimated in dense forests (Figure 3-8a). This could be explained by their different processing mechanisms. In dry dipterocarp forests that have low fractional cover in the dry season, the light-colored stems/branches and senescent leaves are more easily mis-classified as bare soil in an IKONOS image. However, they are important non-sky components in the hemispherical photos. In dense evergreen forests, the in-tree gaps less than 1x1 1n were not observable in the IKONOS image, whereas they played an important role in hemispherical photos. The small-area 55 gaps in the IKONOS classification images were also smoothed out during the majority filter processing. These differences aside, Figure 3-8a shows that the IKONOS fc is highly correlated to ground measurements (R2 =0.97). This indicates that where intensive ground measurements are lacking, the IKONOS fc could serve as ground truth for the purpose of validation. The corresponding ETM+fc values at the limited study sites were also plotted in Figure 3-8a. They too are highly correlated with the IKONOS fc (R2 =0.96). The correlation line is close to the 1:1 line, indicating that the ETM+ image can be processed to estimate fc with high accuracy. Instead of the limited ground measurements, the IKONOS estimated fc values could be compared with the ETM+ estimation in any corresponding areas. In this study, four subset areas in different forest types were chosen for the comparison: dry dipterocarps, mixed deciduous, transition zone, and evergreen forests. Since it is impossible to identify pine transition forests in the ETM+ image, the transition zone covers mixed deciduous, pine transition, and dry evergreen forests. The subset of evergreen forests covers both dry evergreen and moist evergreen forests because it is difficult to identify them in the ETM+ image. The subset of evergreen forests was chosen nearby the peak of Mount Doi Inthanon. In each subset, 100 positions were randomly selected to avoid the autocorrelation effect. The fc value at each position was an average of a 3x3 window to reduce the possibility of geometric mismatch between the ETM+ and IKONOS images. 56 As shown in the scatterplot (Figure 3-8b), the accuracy of the ETM+fc estimation varies with different forest types. For dry dipterocarps that have a low forest density, the fc values of both the IKONOS and ETM+ estimation are more scattered than other forest types. Most of the ETM+fc values are clustered in a range of 20% to 40% while the IKONOSfc values ranges from 10% to 60%. The distribution of the ETM+fc in dry dipterocarp forests is smoother than that of the IKONOS, but the values are underestimated. The ETM+ estimation is better for the mixed deciduous forests whose density is higher. However, the ETM+fc is still underestimated, ranging between 30% and 60%, whereas the IKONOS fc could be as high as 80%. For the transition zone that has a much higher forest density, the ETM+ and IKONOS fc match well. Both the ETM+ and IKONOSfc values range from 60% - 90%, and the data points are closer to the 1:1 line. For the moist evergreen forests where the density is very high, both ETM+ and IKONOS estimated fc values are saturated. All points in Figure 3-8b are more or less scattered along the 1:1 line. The ETM+fc is correlated with the IKONOS fc R2: 0.70). Similar to Figure 3-6a, it indicates that the ETM+ image is a good source of forest fractional cover estimation over large areas. Figure 3-8b also shows that when forest cover is not saturated, the higher the forest density, the higher the accuracy of the fc estimation. Figure 3-9 presents the H(ONOS image, the corresponding subset of the ETM+ image and their fc maps. To avoid the seasonal variation, only the IKONOS image acquired in the dry season (February 27, 2000) was compared. The IKONOS and ETM+ estimated fc distributions show both similarities and differences in each forest type. In the fallowed 57 agriculture fields in the lower left of the subset, the IKONOS estimated fc values are around 0, and the ETM+ estimated fc values are less than 20%. In dry dipterocarp forests in the upper left of the subset, both the IKONOS fc and ETM+fc have a wide range of distribution. However, the IKONOS fc is more scattered with a range of 0-60%, while the ETM+fc has a smaller range of 20-40%. Conversely, in the evergreen forests in the right of the subset, the IKONOS fc distribution is less variable than that of the ETM+. In the mixed deciduous forests in the middle of the subscene, the fc distributions from the two images are very similar. The secondary forests that are visually interpreted in both images can also be identified in the fc maps. In the middle of the subset, there is large area of regeneration of evergreen forests in their early stage. Both ETM+ and IKONOS fc values are in a range of 20-60%, much lower than the natural, mature evergreen forests nearby. In the late-stage secondary forests such as the isolated small areas in the lower right of the subset, the ETM+fc map has fractional cover values around 60-80%. However, in the IKONOS fc map, the values exceed 90% and cannot be separated from the dense, mature evergreen forest nearby. The differences between the IKONOS and ETM+ estimated fc values are mostly at local scales. When all forest types are considered in the whole subset, the ETM+ estimated fc distribution shows high similarity with the one from the IKONOS image. In Figure 3-9, from west to east of the subset, the forest types change from dry dipterocarps, mixed deciduous, and evergreen forests as visually interpreted in the ETM+ image. The fc 58 values in both the IKONOS and ETM+fc maps also show a gradual increase from 0 to 100%. As seen in Figure 3—8 and 3-9, the accuracy of ETM+fc estimation is high when compared with the IKONOS fc that is presumed correct. The accuracy also varies with forest types in different fractional cover densities. For dry dipterocarps that have low fractional cover in the dry season, the fc is underestimated with the ETM+ image, although its distribution is less variable than the IKONOS fc. For mixed. deciduous and transition zone forests, the correlation between the ETM+ and IKONOS estimated fc values are the highest among all forest types. For evergreen forests, especially moist evergreen forests, both the ETM+ and IKONOS estimated fc tends to saturate. The ETM+ estimated fc map also reveals the different stages of forest clear-cut and regrowth in evergreen forests. In general, the higher the forest density, the higher the accuracy of ETM+fc estimation. Most of the tropical forests have high density and, therefore, the method presented in this study is promising for large-area fractional cover estimation, tropical forest evaluation, and ecosystem management. 3.5.3 Seasonal adjustment of fractional cover The green fractional cover in the wet season is a more realistic representation of the fractional cover distribution when all types of forests are leaf-on. However, sky conditions during the wet season in the tropics are often cloudy or hazy. For satellite sensors with coarse temporal resolution and wide swath such as ETM+ (16 days, 180km swath width), it is often difficult to acquire clear images during the wet season. It is, 59 however, possible for the IKONOS satellite who has a sensor with a more flexible repeat period (an average of 1.5 days). The IKONOS images acquired in the wet season can be used to adjust the ETM+fc from the dry season to the “true” green cover of the peak growing season. Nine IKONOS images were acquired on October 9th, 2002 (wet season). These images covered 28 study sites, among which 7 sites were measured in August 2001 (wet season). One IKONOS scene was acquired on February 27th, 2000 (dry season), covering 7 study sites, among which 6 sites were measured in January 2002 (dry season). Figure 3-10 shows the relationships between the IKONOS and ETM+fc in wet and dry seasons. Three series of data points are displayed: IKONOS and ETM+fc values in the dry season (7 points), IKONOSfc in the wet season, but ETM+fc in the dry season (28 points), and ground-measuredfc in the wet season, but ETM+fc in the dry season (12 points). The 7 points of IKONOS vs. ETM+fc in the dry season are closely distributed along the 1:1 line, indicating again that the ETM+fc during the dry season is estimated with high accuracy. The relationships of IKONOS + ground fc in the wet season vs. ETM+fc in the dry season, however, is not linear. Due to the seasonal variation of forest greenness, the fc in the wet season, especially in dry dipterocarps and mixed deciduous forests, is much higher than the fc calculated during the dry season. The fc in the pine transition and evergreen forests does not change much across the seasons. Both IKONOS and ground- measured fc values in the wet season follow a similar trend and, therefore, can be combined as ground truth. 60 As shown in Figure 3-10, the ground truth (H(ONOS and ground-measuredfc) from the wet season is logarithmically related to the ETM+fc calculated from the dry season. A simple logarithmic regression was processed and the regression line is shown in Figure 3- 10. With the obvious three outliers in the upper left of the figure, the expected value is higher than the observed when fc is less than 50%. The squared correlation coefficient (R2) of the regression is only 0.64. When the three outliers were removed, a logarithmically curve fitting was also made with the following equation: fcwe, = 0.30 * ln(fcd0, /100 + 0.09) + 0.91 when fc < 91% 3-11 few, 2 fem when fc Z 91% ( ) Here, fc is the expected fc in the wet season, and fem, is the observed fc from the we! ETM+ image in the dry season. Using Eq.3-1 l, the expected fc is only 94% when we! ETM+fc is saturated (100%), which is less than the ground truth. Therefore, fcwe, is underestimated in very dense forests. Since there is not much difference between fc we! and fc.»). in dense evergreen forests, we set few, = fem, when the ETM+ estimated fc 2 91%. The curve fitting is demonstrated in Figure 3-11. The measured data points were closer to the curve fitting line when the outliers were removed. The squared correlation coefficient (R2) for this curve fitting was 0.87, much higher than the simple logarithmic regression in Figure 3-10. The errors of the three coefficients in Eq.3-11 were much lower than the expected values. A )(2 goodness-of—fit was also done to this curve fitting and the 12 = 0.15. For a 3-parameter curve fitting with 37 samples (40 data points —— 3 outliers), the 61 degree of freedom equals 34 (37 — 3 model coefficients). At 95% confidence level, the standard 1305,34 = 48.60. Since 12 is much less than 15053, , we accept the curve fitting described by Eq.3-l 1. With Eq.3-11, the ETM+ estimated fc in the dry season could be adjusted to the “true” green forest fractional cover of the wet season, a more realistic biophysical attribute in tropical forest ecosystems. Figure 3-12 is the adjusted fractional cover map. It is obvious that the distribution is much less spatially variable than the one in the dry season (Figure 3-5). For dry dipterocarps and mixed deciduous forests, the green cover is much higher than the cover calculated from the dry season. The fractional cover of evergreen forests does not change much. Except for the non-forest areas like agriculture fields and settlements whose cover is less than 20%, the green forest fractional cover in the watershed ranges from 20% to 100%. The fractional cover of most forest areas in the watershed is higher than 50% during the growing season. 3.6 Conclusion and Discussion Forest fractional cover is a good indicator of forest integrity. Quantifying the fractional cover in tropical forests is very important for evaluating deforestation and recovery. In this study, a linear unmixing model in a vegetation index domain was built to estimate the forest fractional cover distribution in the tropical forests of the Mae Chaem Watershed in northern Thailand. The Modified Soil Adjusted Vegetation Index (MSA V1), which was most linearly related with vegetation abundance, was chosen as the independent variable in the model. Two endmembers, full-cover forest canopy and open area with bare soil or 62 senescent grass, were calculated through statistical analysis. With the model and the ETM+ image acquired in the dry season, the forest fractional cover distribution in the watershed was mapped. The forest fractional cover distribution in the study area results from the different extent of both human disturbance and seasonal variation. Along the elevation gradient in the mountainous watershed, the forest types change from dry dipterocarps at low elevation, to mixed deciduous, pine transition, dry evergreen, and finally moist evergreen forests on the top of the Mount Doi Inthanon. The forest fractional cover increases along this gradient (downhill to uphill) in a similar trend to the forest type change. The dry dipterocarp forests are frequently burned for agriculture and regrow after the agriculture fields are abandoned. Moreover, as deciduous forests, the dry dipterocarps vary seasonally. During the dry season, the leaves of dry dipterocarp forests become brown and fall off. Therefore, as estimated with the ETM+ image, the dry dipterocarps have a lower fractional cover between 10-40% in the dry season. The mixed deciduous forests are composed of both deciduous and evergreen species. Also, because of the relative difficulty to access these types, the mixed deciduous forests are less disturbed by human activities. As a result, their fractional cover is between 40-80% in the dry season. The pine transition and evergreen forests are composed of only evergreen species. Similarly inaccessible due to the high topography, they are not degraded much as the dry dipterocarps and mixed deciduous forests. The fractional cover of these forests is generally higher than 70% in the dry season, and the seasonal variation is minimal. For 63 the moist evergreen forests on the top of Mount Doi Inthanon, the fractional cover is often saturated (i.e. 100%). In each forest type, the fc distribution reveals different stages of forest degradation and recovery in the study area. For example, there were isolated areas of clear-cutting in the evergreen forests in the early 19905. The regrowth of these areas shows in the fractional cover map with the cover varying between 20 to 80%, lower than the surrounding mature evergreen forests. The forest fractional cover estimated with the ETM+ image was validated by isolated ground measurements in the watershed. The ground-measuredfc was calculated using the GLA software to process hemispherical photographs taken on the ground with a fisheye lens. The ETM+fc values were correlated with those at the 32 study sites measured in the dry season with R2: 0.76. Similar to the trend of ETM+fc, the ground-measuredfc increased with elevation as the forest types changing from dry dipterocarps to mixed deciduous, pine transition, and finally dry and moist evergreen forests. Due to the methodological difference, the ETM+fc was lower than the ground—measuredfc in low- density forests such as dry dipterocarps, and higher than ground-measuredfc in dense forests such as the tropical evergreen. The ETM+fc was saturated in moist evergreen forests at high elevations. The forest fractional cover estimated with the ETM+ image was further validated in a large spatial area by one hi gh-resolution IKONOS image that was also acquired in the dry 64 season, 25 days after the acquisition of the ETM+ image. The IKONOS fc was calculated by image classification and visual interpretation. With limited study sites covered by this scene (7 sites only), the squared correlation coefficient between the ground-measuredfc and the IKONOS fc was 0.97, indicating the high probability of the IKONOS image serving as an accurate validation source in this study. Also at these study sites, the ETM+ fc was highly correlated with the IKONOSfc with R2: 0.96. The correlation line was very close to the 1:1 line in the scatterplot. The ETM+fc was also compared with the IKONOS derived fc in a larger dataset, in which 100 points in each forest type were randomly selected for the purpose of comparison (in total of 400 points). In the ETM+ vs. IKONOSfc scatter plot, these 400 points were loosely clustered along the 1:1 line. The squared correlation coefficient was 0.70. The ETM+fc estimated in the dry season can be adjusted to the “true” green cover with ground measurements and IKONOS images from the wet season. A logarithmical relationship was observed and a curve fitting model was built using the combined ground-measured and IKONOS fc in the wet season and the ETM+fc in the dry season. With this relationship, the ETM+fc map in the dry season was adjusted to wet season conditions, a better representation of tropical forest canopy cover. The successful estimation of forest fractional cover is an obvious improvement over simple forest/non—forest binary classifications. Knowing the fractional cover distribution in tropical forests provides a better way to measure both forest degradation and its 65 recovery. It is also very helpful for estimating forest biomass in tropical forests and understanding their role in carbon sequestration and global change. 66 3.7 References Adams, J. B., Smith, M. O. and Gillespie, A. R. (1993). Imaging spectroscopy: interpretation based on spectral mixture analysis. 145-166. In: Remote Geochemical Analysis: Elemental and Mineralogical Composition. Pieters, C. M. and Englert, P.A. (Eds), Cambridge University Press, Cambridge. Barros, A. C. and Uhl, C. (1995). Logging along the Amazon River and estuary: patterns, problems, and potential. Forest ecology apd Management: 77, 87-105. Carlson, T. N., Perry, E. M. and Schmugge, T. J. (1990). Remote estimates of soil moisture availability and fractional vegetation cover for agricultural fields. Agriculture, Forestry, and Meteorology: 52, 45-70. Carlson, T. N. and Ripley, DA. (1997). On the relation between NDVI, fractional vegetation cover, and leaf area index. Remote Sensing of Environment: 62, 241- 252. Cochrane, M. A., Alencar, A., Schulze, M. D., Souza Jr., C. M., Nepstad, D. C., Lefebvre, P. and Davidon, E. A. (1999). Positive feedbacks in the fire dynamic of closed canopy tropical forests. Science: 284,1833-1835. DeFries, R. S., Hansen, M. C., Townshend, J. R. G., J anetos, A. C. and Lovelands, T. T. (2000). A new global l-km dataset of percentage tree cover derived from remote sensing. Global change biology: 6, 247-254. Frazer, G. W., Canham, C. D. and Lertzman, K. P. (1999). Gap Light Analyzer (GLA), Version 2.0: Imaging software to extract canopy structure and gap light transmission indices from true-color fisheye photographs, users manual and program documentation. Copyright © 1999: Simon Fraser University, Burnaby, British Columbia, Canada, and the Institute of Ecosystem Studies, Millbrook, New York, USA. Gobron, N., Pinty, B., Verstraete, M. and Govaerts Y. (1999). The MERIS Global Vegetation Index (MGVI): description and preliminary application. lntemational Journal of Remote Sensing: 20 (9), 1917-1927. Gutman G. and Ignatov, A. (1998). The derivation of the green vegetation fraction from NOAA/AVHRR data for use in ntunerical weather prediction models. International Joumal of Remote Sensing: 19(8), 1533-1543. Hanan, N. P. and Prince, S. D. (1991). Spectral modeling of multicomponent landscapes in the Sahel. lntemational Journal of Remote Sensing: 12 (6), 1243-1258. Huete AR. (1988). A soil-adjusted vegetation index (SAVI). Remote Sensing of Environment: 25, 295-309. 67 Huete, A.R., Justice, C. and Van Leeuwen, W. (1999). MODIS vegetation index, MODIS algorithm theoretical basis document, NASA Goddard Space Flight Center, Greenbelt. Laurance, W. F ., Laurance, S. G., F erreira, L. V., Rankin-de Merona, J ., Gascon, C. and Lovejoy, T. E. (1997). Biomass collapse in Amazonian forest fragments. Science: 278,1117-1118. J asinski, M. F. (1990). Sensitivity of the normalized difference vegetation index to subpixel fractional cover, soil albedo, and pixel scale. Remote Sensing of Environment: 32, 169-187. J asinski, M. and Eagleson, P. S. (1990). Estimation of subpixel vegetation cover using red-infrared scattergrams. IEEE Transactions on Geoscience and Remote Sensing: 28 (2), 253-267. Myneni, R. B., Asrar, G., Tanre, D. and Choudhury, B. J. (1992). Remote sensing of solar radiation absorbed and reflected by vegetated land surfaces. IEEE Transactions on Geoscience and Remote Sensing: 30, 303-314. Myneni, R. B. and Williams, D. L. (1994). On the relationship between FPAR and NDVI. Remote Sensing of Environment: 49, 200-211. Pinty, B. and Verstraete, M. M. (1992). GEMI: a non-linear index to monitor global vegetation from satellites. Vegetatio: 101, 15-20. Price, J.C. ( 1992). Estimating vegetation amount form visible and near infrared reflectances. Remote Sensing of Environment: 41, 29-34. Qi, J ., Chehbouni, A, Huete, A. R. and Kerr, Y. (1994). A modified soil adjusted vegetation index (MSAVI). Remote Sensinggf Environment: 48, 119-126. Qi, J ., Marsett, R. C., Moran, M. S., Goodrich, D. C., Heilman, P., Kerr, Y. H., Dedieu, G., Chehbouni, A. and Zhang, X. X. (2000). Spatial and temporal dynamics of vegetation in the San Pedro River basin area. Agricultural and Forest Meteorology: 105, 55-68. Quarmby, N. A, Townshend, J. R. G., Settle, J. J ., Milnes, M., Hindle, T. L., and Silleos, N. (1992). Linear mixture modeling applied to AVHRR data for crop area estimation. International Journal of Remote Sensing: 13, 415-425. Radeloff V. C., Mladenoff, D. J. and Boyce, M. S. (1999). Detecting Jack Pine budworm defoliation using spectral mixture analysis: separating effects from determinants. Remote Sensing of Environment: 69, 156-169. Rahman, H., Pinty, B. and Verstraete, M. M. (1993). Coupled surface-atmosphere reflectance (CSAR) model, 2, semi-empirical surface model usable with NOAA 68 advanced very high resolution radiometer data, Journal of Geophysical Research: 98(20), 781-801. Settle, J. J. and Drake, N. A. (1993). Linear mixing and estimation of ground cover proportions. lntemational Journal of Remote Sensing: 14, 1159-1177. Souza Jr. C. and Barreto, P. (2000). An alternative approach for detecting and monitoring selectively logged forests in the Amazon. lntemational Journal of Remote Sensing: 21, 173-179. Stone, T. and Lefebvre P. (1998). Using multi-temporal satellite data to evaluate selective logging in Para, Brazil. International Journal of Remote Sensing: 19, 2517-2526. Thailand Land Use and Land Cover Change Case Study (1997). Southeast Asia, IHDP, IGBP, WCRP program. Townshend, J. R. G. (1999). MODIS enhanced land cover and land cover change product: algorithm theoretical basis documents (ATBD), version 2.0. Tucker, J. J. (1979). Red and photographic infrared linear combinations for monitoring vegetation. Remote Sensing of Environment: 8, 127-150. Verhoef, W. (1984). Light scattering by leaf layers with application to canopy reflectance modeling: the SAIL model. Remote Sensing of Environment: 16, 125-141. Wessman, C. A, Bateson, C. A. and Benning, T. L. (1997). Detecting fire and grazing patterns in tallgrass prairie using spectral mixture analysis. Ecological Applications: 7, 493-511. Wittich, K. P. and Hansing, O. (1995). Area-averaged vegetative cover fraction estimated from satellite data. lntemational J oumgl of Biometeorology: 38, 209-215. Wittich, K. P. (1997). Some simple relationships between land-surface emissivity, greenness and the plant cover fraction for use in satellite remote sensing. International Journal of Biometeorology: 41, 58-64. Zeng, X., Dickinson, R.E., Walker, A. and Shaikh M. (2000). Derivation and evaluation of global l-km fractional vegetation cover data for land modeling. Journal of Applied Meteorology: 39(6), 826-839. 69 Mae Chaem l 02/27/2000 (a) (b) Figure 3-1 ETM+(band4+3+2, 02/02/2000) (a) and IKONS images (band4+3+2, 02/27/2000 and 10/09/2002) (b) in Mae Chaem Watershed. The dots in (b) represent the study sites in two field trips. 70 - 1707-1991 [:1 1992— 275 [:1 2276— 2500 (a) (b) Figure 3-2 DEM data with hillshade effect (a) and topographic correction with Rahman’s BRDF model (b). 1.0 - >< _ g 0'8 _._ GEMI E 0 6 + NDVI g ‘ + SAVI "g —— MSAVI ‘5 0'4 ‘ + EVI 9 + MGVI > 0.2 I 0.0 I l I l o 1 2 3 4 5 . Leaf Area Index Figure 3 -3 Dynamic ranges of the six vegetation indices calculated with simulated spectral data in SAIL model. 71 Figure 3-4 MSAVI image in Mac Chaem Watershed. -I - Figure 3-5 Forest fiactional cover map in Mac Chaem Watershed. 72 ETM+ estimated fc (%) (02l02/2000) 0 I I I I 0 20 40 60 80 100 ground measured fc (%) (01/I20-2712002) (a) 100-» - 80 I .3 aground .2 I, a n ETM+ . I59. fine dry dipter. nixed pine dry ever. rroist decid. trans. ever. 0)) Figure 3-6 Comparison of ground-measured and ETM+ estimated fc: all forest types combined (a) and by forest type (b). 73 200 a 150 ~ 2 E wt g 100 - .o q. o g o forest(bright+dark) a bare 50 ~ A brightopen + dark open x shaded bare 0 water 0 l T l l O 50 100 150 200 DN of band 3 (Red) Figure 3-7 Feature space of different classes in the supervised classification of IKONOS images. 74 100 [ I I l . l ' /$ 1 00 Z i ' / —e—fc_ground_Jan : j j CG, / Q 90 “ Y=M0+M1‘X ’(‘cu‘ M0 32.38| ; [1‘1 3 80 M1 osozoelmgm _ g 80 Z S R 0.98309' + I I I - —h In 5 f ____________ i ...... ; __ O 8 7° .‘ s 2 2 , 5 7° 5' In . . : s ' o. .0 60 , 60 m .E I I C / I . . g s i <1: F . . ’ .9 50 — --------- ------- . ---------- / -------- g ~~~~~~~ g IIIIIIIIII i ----------- 1 ---------- — 50 ‘8 1:3 3 . ,’ 3 3 a = 9 3 3 a ,/ 3 I .3 -I{}>—fc_ETM_Feb 11 U) f ‘ 3 ; 5 Y=M0+M1'X S ; ‘3" I . ; || M0 0.37416] 30 7' "' 3' " ' ‘ g‘ "l M1 1.o1o1|[ 30 I i II R 0.9804611 20 l .__ l l l l I 20 2o 30 40 50 60 70 80 90 100 IKONOS fc in dry season (Feb) (a) 100 0 dry dipterocarps 00 a: mixed deciduous ’3‘ 8O - otransition zone g) Q E X o 0% A I r: evergreen q) o 3 O x o 8 60 ~ 0 g '” 6o 5 x 8 4O - + E *° In 20 - R2: 0.70 O I I I l O 20 40 60 80 100 IKONOS fc (dry season) (%) (b) Figure 3-8 Scatter plots of ground-measured and IKONOS estimated fc in 6 sites (a), and the ETM+ and H(ONOS estimated fc in 400 randomly selected points (b). 75 , ETM+ image (4+3+2), 02/02/00 IKONOS 1mage(4+3+2), 02/27/00 IKONOS estimated fc ETM+ estimated fc Figure 3-9 IKONOS and ETM+ subset images and their fc maps. 76 100 g 80 — .2 8 z 60 - O E '2 a 40 — 1: § 2 20 ,x” Ofc_ground_wet m I x” 2 X fc__ikonos_wet // R ”0'64 Ifc_ikonos_dry O I I I I I O 20 4O 60 80 1 00 ETM+ to in dry season (%) Figure 3-10 The ground-measured, H(ONOS and ETM+ estimated fc scatterplot in dry and the wet seasons. c: 0.8 r O U) (U Q) m . 5 0.6 - 3 .E 8 E 0.4 - . ‘ ~ - 8 y = m1 + m2 * In(M0+m3) §3 Value Error m m1 0.907 0.017 C2) 0.2 m2 0.298 0.049 — 0 m3 0.086 0.053 5 Chisq 0.145 NA R 0934 NA 0 4 l 0 0.2 0.4 0.6 0.8 1 ETM+ fc in dry season Figure 3-11 The curve fitting model to adjust fc in the dry season to the wet season. 77 . ._-.I A ‘5 u ‘II-IUIM Figure 3-12 Forest fractional cover map adjusted from the dry season to the wet season. 78 Chapter 4 Estimation of Leaf Area Index with Fractional Cover Data in Tropical Forests 4.1 Introduction Leaf area index (LAI), the total single-sided leaf area per unit ground area, is another important biophysical attribute of forest ecosystems. It controls the partitioning of incoming solar energy into sensible and latent heat fluxes and is a good indicator of energy, gas and water exchanges between land surface and atmosphere. Because of its important rule in the physiological processes such as photosynthesis, transpiration and evapotranspiration, LAI has been widely used to parameterize and validate models of ecosystem functioning, biosphere-atmosphere interaction, vegetation growth, net primary production and other environmental processes at landscape to global scales (Sellers and Schimel 1993). The direct measurement of LA] involves cutting and measuring the leaves in a small area. It provides the most accurate measurements. However, this method is very time consuming and destructive. As a result, this method is not recommended for long-term large scale measurements. An alternative way is to indirectly measure LA] with commercial instruments like the Li-Cor LAI-2000 and fisheye photography (Welles 1990). The accuracy of these indirect methods highly depends on the sunlight radiation model that these instruments use and the atmosphere conditions at the time of 79 measurement. Both direct and indirect LAI measurements provide ground data for the validation of remote sensing methods at large scales. One of the primary approaches to estimate LAI using remote sensing imagery is the polynomial regression between LAI and vegetation indices (VIs) from radiometric measurements (Qi et al. 2000; Baret 1995; Best and Harlan 1985; Peterson et al. 1987). This relationship is non-linear (Tucker et al. 1981) and saturates at LAI 2 3 depending on the type of vegetation index used, the canopy type studied, and the experimental conditions (Spanner et al. 1994; Baret and Guyot 1991). This empirical LAI— VI relationship is also affected by leaf pigments, internal structure, orientation distribution, leaf clumping, woody reflectance, and heterogeneity in tree height and tree gaps (Turner et al. 1999). Another commonly applied approach is to estimate LA] by the inversion of bi-directional reflectance distribution function (BRDF) models (Pinty et al. 1990; Goel and Thompson l984a,b; Qi et al. 1995). So far, more than a dozen BRDF models have been developed for various types of surface such as cropland, grassland, and bare soil surfaces (Strahler 1994). Some of these models are mathematically invertible so that biophysical attributes can be calculated. However, a BRDF model often has multiple input parameters. For large-scale model inversion, some of the essential input forest parameters are often unavailable and have to be simplified. This simplification often results in unreasonable LAI values for forests with high heterogeneity of distribution and topography. 80 Because of the topographical difficulty to access, it is impossible to measure LA] with LAI-2000 instrument at study sites in the watershed. However, as discussed in Chapter 3, the forest fractional cover was measured with a fisheye camera at each study site, and the forest fractional cover product from remotely sensed data was validated. In this chapter, a simple regression model was developed to retrieve forest LAI distribution in the study area from forest fractional cover product in Chapter 2. The adjustment of LA] product from the dry season to the wet season was also performed. 4.2 Experimental Design and Field Measurements 4.2.1 LAI-2000 and fisheye photographs A variety of instruments have been developed which apply radiation models to indirectly measure the vegetative canopy attributes. When passing through the canopy, solar radiation is mainly alternated by foliage amount and orientation. The spatial distribution of the foliage is assumed to be random in the radiation model. The common strategy of the radiation model is to describe how radiation is affected as it passes through a canopy with some well-defined, geometrical canopy attributes, then make appropriate radiation measurements and invert the model to estimate the value of these attributes (Welles and Cohen 1996). The success of the radiation model is thus determined by how closely the real canopy conforms to the idealized one. It should be noted that, when passing through a canopy, the radiation is not only affected by leaves, but any other opaque object like stems, fruits, and branches. As a consequence, 81 although LA] is defined as leaf area index, the LA] estimates with indirect methods should be better described as plant projected area index (Welles and Norman 1991). The LAI-2000 Plant Canopy Analyzer developed by LI-COR company is the most commonly used commercial instrument to measure leaf area index and some other vegetative canopy attributes. It uses hemispherical optics and a ringed detector that simultaneously measures diffiise radiation in five distinct angular bands about the zenith (Welles and Cohen 1996). For a single-sensor LAI-2000 equipment, a reference reading is made at the open area above the canopy, instantly followed by one or more below canopy readings. Sometimes another open reading is made after the below canopy readings. For each ring, the canopy’s gap fraction at the view angle of that ring is assumed to be the ratio of the below and above canopy readings (Welles and Norman, 1991). With a sunlight radiative transfer model, the LAI, foliage amount density, and foliage orientation can be calculated. The LAI-2000 has proven to be very useful in short canopies like agricultural crops and grasses. However, for tall canopies like forests, the application of LAI-2000 is limited. For each LAI measurement, a simultaneous reference reading is needed in an open area that is large enough to avoid the effects of forest scattering. In this study, all of the study sites are located in the deep forests in Mae Chaem Watershed and, therefore, it was impossible to measure LA] with the LAI-2000 instrument during the two field trips. 82 As discussed in Chapter 3, the digital photos with fisheye lens were taken at the study sites and processed with the Gap Light Analyzer (GLA) software to calculate the gap fraction. A similar radiation model was also applied in GLA to calculate LAI over certain zenith angles. For example, LAI-4ring is the effective leaf area index integrated over the zenith angles 0° to 60°, and LAI-5ring is the effective leaf area index integrated over the zenith angle 0° to 75° (Frazer et al. 1999). The forest fraction (in 0-1 scale) is 1.0 — gap fraction. As shown in Figure 4-1, LA] values at all study sites from the two field trips are exponentially related to the forest fractional cover. In sparse forests, there is more open space at larger zenith angle (60°-75°) and, therefore, the values of LAI-5ring are lower than LAI-4ring. Contrarily, in dense forests, the fractional cover is higher than gap fraction and therefore, the values of LAI-5ring are higher than LAI-4ring. Similar to the LAI-vegetation index relationships, the LAI-forest fractional cover curve begins to saturate at LAI around 3.0. However, the LA] values measured with fisheye photos are unreasonably low in Figure 4-1. Even for the dense moist evergreen forests with fractional cover higher than 95%, the LAI values are less than 4.0. The radiation model in GLA software assumes that the atmosphere conditions above the forest canopies are constant at the time of measurements at all of the study sites, which in fact varies greatly depending on different seasons, amount of cloud cover or cloud properties. It could be the possible reasons for the low LAI values in Figure 4-1. 83 4.2.2 LAI ground measurements From the radiation model in GLA software, the LAI and fractional cover are logarithmically correlated (Figure 4-1). However, the LA] values in GLA software with fisheye photos are underestimated. While the forest fractional cover can be measured with fisheye photos, LAI-2000 provides a good measurement of LA]. Since it is impossible to use LAI-2000 in Mae Chaem Watershed, I made additional measurements during the third field trip in the hardwood forests in northern Michigan in October and November 2002. Between October 1 and November 12, 2002, four forested areas were visited and totally 42 study sites were measured (Table 4-1). These sites were close to river, lakes, or large open areas so that the LAI-2000 requirements were met. At each study site, two open readings and five under-canopy readings were recorded with LAI-2000 to measure LAI, following the TIIIII sequence. Three digital photos were taken with the fisheye lens and were processed using GLA software to calculate fractional cover and LAI. As shown in Figure 4-1, for forests with fractional cover lower than 95%, the LAI-4ring and LAI- 5ring values do not change much when calculated in GLA software. Therefore, only LAI- 4ring readings are compared in this analysis. The study sites visited in early October simulates the canopy conditions of tropical forests with higher cover, such as mixed deciduous and tropical evergreen in Mae Chaem Watershed (in the dry season). The forests in northern Michigan are typical deciduous broadleaf forests. In late October and early November, the leaves become yellowish and 84 begin to drop off which is analogous for canopy conditions of dry dipterocarps and sparse mixed deciduous forests (in the dry season) in Mae Chaem Watershed. Therefore, all these study sites at different dates could approximately represent the ground measurements in our study area in Mae Chaem Watershed and may provide ground data for the LA] estimation. It should be noted that, the structure of tropical forests and hardwood temperate forests are different. Some errors may be introduced with this approximation. The LA] values from LAI-2000 and fisheye photos over study sites in northern Michigan were compared in Figure 4-2. The 1:1 line was also drawn in the figure to demonstrate the variation of the two LAI measurements. Although LAI values from fisheye photos are a little higher than the ones from LAI-2000 for forests at low fractional cover, they are much lower for forests with higher fractional cover. The LA] values from fisheye photos increase very slowly when forest cover is higher. Among all the study sites with different dates (leaf-on and leaf-off), the LAI-2000 measurements range from 0.88 to 5.04, while the LAI values from fisheye photos (LAI 4-ring) only change from 1.39 to 2.66. Figure 4- 2 confirms that fisheye photos are not a good source of LAI measurements. 4.2.3 LAI ~ forest fractional cover relationship As discussed in Chapter 2 and 3, fisheye photos provide reasonable forest fractional cover measurements. Replacing LAI values in Figure 4-1 with the LAI-2000 measurements, the relationship between LA] and forest fractional cover was shown in Figure 4-3. Most forests in Michigan are old secondary or natural forests and there was 85 no sparse forests found at the study sites. Therefore, the forest fractional cover values at all study sites in Figure 4-3 are higher than 0.6. The LAI and forest fractional cover values fit an exponential relationship. 4.3 Model Development and Results 4.3.1 A modified Gaussian regression model A modified Gaussian curve fitting model was developed with the ground measurements in northern Michigan. In Figure 4-4a, both the LAI and forest fractional cover values are extended to the origin to represent the sparse forests with fractional cover from 0 to 0.6, a simulation of dry dipterocarp forests in the dry season in Mae Chaem Watershed in Thailand. The residuals of the regression are plotted in Figure 4-4b. Among the 42 study sites, there are only four sites with absolute residual values larger than 1.0, and all these 4 sites have fractional cover around 0.85 or higher. Most of the residuals are distributed between i0.5. Figure 4-4 also indicates that the higher the forest fractional cover, the higher the residuals. It is reasonable because the LAI-2000 measured LA] is a measure of plant projected area instead of green foliage area. As the fractional cover becomes higher, there introduces more error in the LAI—2000 readings. The regression equations are expressed as: LAI = 0.217 + 0.058 * e‘5’578‘If"°-°33”> 0.6 < fc $1.0 LAI = 0.85 * fc 0.0 < fc s 0.6 (4-1) Here fc is the forest fractional cover at the scale of [0,1]. There was no ground data when fc is lower than 0.6. However, from Figure 4-2, the relationship between LAI and 86 fractional cover is almost linear when the fractional cover is low. Therefore, a linear relationship in Eq.4-1 was used for 0.0 < fc S 0.6 . A x2 goodness-of-fit test was examined to the modified Gaussian regression. The x2 was calculated from: N (LAIO, —LA1,, )2 12 =2 bLAI p (4-2) i=1 exp Here N is the total number of ground measurements, N=42 in the study sites in the northern Michigan. LAlob is LAI-2000 measurements, LAI,Xp is the modeled LAI values S from Eq.4-1. At the degree of freedom of 38 (N-4 coefficients in the model) and confidence level of 95%, the critical value, 130,38 , is 55.76. As shown in Figure 4-4a, the x2 value in this test is 9.712, far less than 130538 . Therefore, the null hypothesis cannot be rejected and the modified Gaussian regression model is valid. Moreover, the correlation coefficient (R) between modeled and measured LAI values is 0.895, also indicating that the model could characterize the LA] ~ forest fractional cover relationship very well. 4.3.2 LAI estimation The forest fractional cover distribution in Mae Chaem Watershed in Thailand has been mapped in chapter 3. With the modified Gaussian regression model, the LAI distribution could be mapped. Figure 4-5 is the LA] distribution in the study area in dry and the wet seasons, resulting from the forest fractional cover distribution in both seasons. The open areas that are non-forests are masked out. 87 Similar to that of forest fractional cover in Chapter 3, the seasonal change of LA] distribution is significant. In the dry season, the leaves on deciduous trees become senescent and were partially off and, therefore, most of the dry dipterocarps and mixed deciduous forests have very low LAI values (less than 1.0). The tropical evergreen forests have LAI values 4.0 or higher. In moist evergreen forests, the LA] values saturate at around 10.9. In Figure 4-5a, there is a very narrow zone between mixed deciduous and tropical evergreen forests where the LA] values are around 1.0 (corresponding to the fractional cover 70-80%). It could be the pine zone transition where the seasonal change is less significant than deciduous or mixed deciduous forests, but more significant than tropical evergreen forests. In the wet season, the tree leaves are green and healthy, and most of the dry dipterocarps and mixed deciduous forests have LAI higher than 1.0 (Figure 4-5b). The narrow zone in Figure 4-5 a merges to the tropical evergreen forests with LAI greater than 4.0. The LA] increment in tropical evergreen forests is not as significant as other forest types, but there are more saturated areas than those in the dry season. The open areas in Figure 4-5b are human settlements (town, villages) and agricultural fields. Surrounding these open areas, there are some forests that have LAI values far less than 1.0 in the wet season. Most of these forests are sparse young regrowth of dry dipterocarps and are highly disturbed by human activities. Therefore, the LAI values in these forests are very low even in the wet season. 88 4.4 Validation As discussed earlier, because of the physical difficulty to access the study sites and the lack of open areas in deep forests, it is difficult to measure the leaf area index over the study sites in tropical forests with LAI-2000. As a result, it is impossible to directly validate the modeled LAI distribution mapped in this study. As an alternative, an indirect validation was made in this section. The leaf area index (LAI-4ring) can be calculated from the fisheye photos over the study sites in Mae Chaem Watershed in northern Thailand. As shown in Figure 4-2, although the LAI from fisheye photos were not as sensitive as that of LAI-2000, there was a linear relationship between these two LAI values in the LA] range between 0 and 6. With this linear relationship, the LAI-4ring values at the study sites in the watershed in Thailand could be adjusted to the equivalent values of LAI-2000 measurements. These adjusted LAI measurements were assumed as ground “truth” to validate the modeled LAI distribution in this study. To be consistent with the modeled LAI distribution in the dry season, only the fisheye LAI-4ring values at the 32 study sites during the second field trip were adjusted. The comparison of the modeled and adjusted measured LAI values at the study sites in the watershed was shown in Figure 4-6. In the LAI range between 0 and 4, the modeled LAI fitted with them well and the data were scattered along the 1:1 line. The adjusted LAI was underestimated when the ground LAI was higher. The adjusted LAI was still less than 4 when the modeled LAI reached 6. In moist evergreen forests, the modeled LAI values 89 were saturated at 10.9 as shown in the LA] map, while the adjusted measured LAI values were only around 5. The adjusted measured LAI from fisheye photos was not the real ground truth. Since the fisheye measured LAI was adjusted from that of LAI-2000 measurement when LAI was less than 6.0, it did not provide good validation when the LA] was higher than 6.0 in the watershed. When LAI was less than 6.0, the adjusted measured LAI could be applied for the purpose of validation. The linear regression line between modeled and adjusted measured LAI was very close to the 1:1 line and the square correlation coefficient R2 = 0.67 (Figure 4-7). 4.5 Conclusion and Discussion In this chapter, a modified Gaussian regression model was built to simulate the relationship between LAI and forest fractional cover in the study area. The LAI values were measured from the LAI-2000, and forest fractional cover values were from fisheye photos as discussed in Chapter 3. In the x2 goodness-of-fit test, the actual x2 from the model was only 9.71 , far less than the critical 16.05.33 (55.76) at the 95% confidence level. The correlation coefficient between the modeled and ground-measured LAI values was as high as 0.90. Most of the residuals distributed between i0.5 and only 4 out of the 42 ground measurements had residuals higher than 1.0. All these statistical analyses supported that the model was valid and could be applied to estimate LAI distribution with forest fractional cover product in Chapter 3. 90 The LA] distributions in both dry and wet seasons were mapped in this study. The seasonal change of the LAI distributions was significant. Most of the dry dipterocarps and mixed deciduous forests have LAI values less than 1.0 in the dry season while higher than 1.0 in the wet season. The LAI values of the tropical evergreen forests were always around 4.0 and higher in both seasons, but there were more areas that were saturated because of the forest fractional cover saturation. The LAI values in saturated areas were higher than 10.0. A narrow zone with LAI around 1.0 was shown in the LA] map in the dry season, possibly the pine transition zone whose seasonal variation was lower than deciduous forests but higher than tropical evergreen. On the other hand, some forests around the open areas (villages, agricultural areas) have very low LAI values even in the wet season, indicating intense human disturbances in these forests. The method in this study provided a new approach to map LAI with forest fractional cover which could be estimated from remotely sensed data. It should be noted, however, the ground data were not measured in the study area in Thailand. Because of the topographical difficulty, as well as the limited open area in deep tropical forests, it is impossible to make LAI—2000 measurements in mountainous tropical forests in Mae Chaem Watershed. Instead, some deciduous forests in northern Michigan in different seasons (leaf—on and leaf-off) were selected to approximately represent the tropical forests in the study area. Since the gap fraction is the major key to calculate LAI and forest fractional cover values, this alternative is acceptable if the fractional cover has large range of variation, e. g. from sparse to dense forests at the study sites in northern 91 Michigan. The ground measurements in this study lacked of forest samples with sparse forests that should be considered in the future research. In this study, it was found that although the GLA software also gave the readings of LAI (4-ring or 5-ring), these LAI values are significantly underestimated. LAI-2000 measurements are more realistic than LAI values from GLA with fisheye photos. Theoretically, the GLA/fisheye photo and LAI-2000 apply the same radiation model. The primary difference is that fisheye photo assumes that the atmospheric conditions are constant at all seasons and under all cloud cover properties, whereas the LAI-2000 makes one or two open readings simultaneously with the canopy readings. The gap fraction of a fisheye photo is calculated with binary image clustering techniques. The gap fraction of LAI-2000, however, is the ratio of the canopy readings to the open readings. If this difference can be compensated for the fisheye photo and GLA software, it is possible to make reasonable LAI as well as fractional cover measurements with fisheye photos. Then digital cameras with fisheye lens will become a more economic and efficient tool for ground measurements in tall canopies, especially the dense mountainous forests where LAI-2000 is in limited usage. 92 4.6 Reference Baret, F. (1995). Use of spectral reflectance variation to retrieve canopy biophysical characteristics. In Advances in Environmental Remote Sensing (M. Darson and S. Plummer, Eds. ), John Wiley and Sons, Inc., 34-51. Baret, F. and Guyot, G. (1991). Potentials and limits of vegetation indices for Lai and APAR assessment. Remote Sensing of Environment: 35, 161-173. Best, R. G. and Harlan, J. C. (1985). Spectral estimation of green leaf area index of oats. Remote Sensing of Environment: 17, 27-36. Frazer, G. W., Canham, C. D. and Lertzman, K. P. (1999). Gap Light Analyzer (GLA), Version 2.0: Imaging software to extract canopy structure and gap light transmission indices from true—color fisheye photographs, users manual and program documentation. Copyright © 1999: Simon Fraser University, Burnaby, British Columbia, Canada, and the Institute of Ecosystem Studies, Millbrook, New York, USA. Goel, N. S. and Thompson, R. L. (1984a). Inversion of vegetation canopy reflectance models for estimating agronomic variables. IV: Total inversion of the SAIL model. Remote Sensing of Environment: 15, 237-253. Goel, N. S. and Thompson, R. L. (1984b). Inversion of vegetation canopy reflectance models for estimating agronomic variables. V: Estimation of LAI and average leaf angle using measured canopy reflectances. Remote Sensing of Environment: 16, 69-85. Peterson, D. L., Spanner, M. A, Running, S. W. and Teuber, K. B. (1987). Relationship of thematic mapper simulator data to leaf area index of temperate coniferous forest. Remote Sensing of Environment: 22, 324-341. Pinty, B., Verstraete, M. M. and Dickinson, R. E. (1990). A physical model of the bidirectional reflectance of vegetation canopies, 2: Inversion and validation. Joumalof Geophysical Research: 95(D8), 11,767-11,775. Qi, J ., Cabot, F., Moran, M. S. and Dedieu, G. (1995). Biophysical parameter estimations using multidirectional spectral measurements. Remote Sensing of Environment: 54, 71-83. Qi, J ., Kerr, Y. H., Moran, M. S., Weltz, M., Huete, A. R., Sorooshian, S. and Bryant, R. (2000). Leaf area index estimates using remotely sensed data and BRDF models in a semiarid region. Remote Sensing of Environment: 73: 18-30. Sellers, P.J. and Schimel, D. (1993). Remote sensing of the land biosphere and biochemistry in the EOS era: science priorities, methods and implementation - 93 EOS land biosphere and biogeochemical panels. Global and Planetary Change: 7, 279-297. Spanner, M., Johnson, L. and Miller, J. (1994). Remote sensing of seasonal leaf area index across the Oregon transect. Ecological Applications: 4(2), 258-281. Strahler, A. H. (1994). Vegetation canopy reflectance modeling - recent developments and remote sensing perspectives. Remote Sensing Reviews: 15, 179-194. Tucker, C.J., Holben, B. N., Elgin, J. H. and McMurtrey, E. (1981). Remote sensing of total dry matter accumulation in winter wheat. Remote Sensing of Environment: 1 1, 171 -l 90. Turner, D. R, Cohen, W. B., Kennedy, R. E., F assnacht, K. S. and Briggs, J. M. (1999). Relationships between leaf area index and Landsat TM spectral vegetation indices across three temperate zone sites. Remote Sensingof Environment: 70, 52-68. Welles, J. M. (1990). Some indirect methods of estimating canopy structure. Remote Sensing Reviews: 5(1), 31-43. Welles, J. M. and Norman, J. M. (1991). Instrument for indirect measurement of canopy architecture. Agronomy Journal: 83, 818-825. Welles, J. M. and Cohen, S. (1996). Canopy structure measurement by gap fraction analysis using commercial instrumentation. Journal of Experimental Botany: 47(302), 1335-1342. 94 Table 4-1 Study sites in northern Michigan. GPS Location . Study area Date Study srtes (UT M, WG884) Muskegon River Watershed (552741, 4789298) 10/1/2002 5 Burchfield Park (697576, 4720093) 10/2/2002 5 ' ' 10/5/2002 16 Rose Lake erdlrfe (716891, 4742766) Comm/anon 1 1/12/2002 10 MSU baker (706798,4732483) 1 1/12/2002 6 woodlot 95 x LAI-4ring o LAI-5ring , i I I Leaf Area Index N 823% a. g 0 I . a I 0.2 0.4 0.6 0.8 1 .0 forest canopy cover Figure 4-1 Relationships between LAI (5-ring and 4-ring) with forest fractional cover. The fractional cover is in the range of [0,1]. LAI (Fisheye) OD 0 1 2 3 4 5 6 LAI (LAI-2000) Figure 4-2 Comparison of LAI measurements from LAI-2000 and fisheye photos. The 1:1 line is also drawn in the plot. 96 LAI (LAI2000) 5 ............................... 5 - I g ’8" 4 _ .................................................. .00 “0:10. H O @Oog a 10 %>9 < 3 _ ...................................................... j. ............. A :L O ; O . _ ; OIO O I 3 2 _ , 06m@@QL.—no a 8 C? 1 __.. o .o ........ 9...: ........ 'I O I I I I I 0.65 0.7 0.75 0.8 0.85 0.9 0.95 forest canopy cover (Fisheye) Figure 4-3 LAI ~ forest fractional cover relationships measured with LAI- 2000 and fisheye photos in northern Michigan. 0 0.2 0.4 0.6 forest canopy cover (a) Figure 4-4 0.8 residual 97 1.5 I I f I I I O 1 ... ,.. . 0.. NJ ‘ o o 05 _ ................. I. Q .0 .Q a i 00 o .' O O 0 _ O 8 O . .. .. BQQ . ‘I O 0 c0) 5 I . O 0 @(g): 0 aogo -0.5 - , -- ~ :~ - I -1 s .. .. O o .15 i I I I i 0.65 O 7 0.75 0.8 0.85 0.9 0.95 forest canopy cover (19) A modified Gaussian curve fitting (a) and its residual plot (b). (b) Figure 4-5 LAI distribution in the study area in dry (a) and wet (b) seasons. modeled LAI O I I l I I 0 2 4 6 8 10 12 adjusted fisheye LAI Figure 4-6 Comparison of modeled and adjusted measured LAI over the study sites in the watershed (second trip). 6 O 5 ~ 0 7t o 4 - _l g 3 O . E ° ’ E 2 - . o. o 1 _ o 3: o o R2=0.67 0 1 2 3 4 5 6 adjusted fisheye LAI Figure 4-7 Correlation between modeled and adjusted measured LAI over the study sites with LAI <6. 99 Chapter 5 A Microwave/Optical Synergistic Canopy Scattering Model and its Inversion to Estimate Forest Structure 5.1 Introduction Forest canopy structure is the combination of forest texture (the qualitative and quantitative composition of the vegetation as to different morphological elements) and forest structure (the spatial arrangement of these elements) (Barkman 1979). Forest structural parameters are important input variables in ecosystem functioning and forest harvest models. The understanding of forest structure is also necessary in accurate estimation of forest biomass. Depending on the study interest, the forest canopy structure can be described at several levels of integration: forest/non-forest at large scales, forest types in different communities, within-forest patches (e.g., successional developments phases, leaf area index), individuals (e. g., species and tree size, height), plant parts (e.g., crown and stem), and plant organs aboveground (leaves, branches, flowers, etc) (Bongers, 2001). From this definition, both forest fractional cover and leaf area index are structural parameters. They are described in Chapter 3 and Chapter 4 because they are all green leaf properties and have been intensively studied using optical remote'sensing. The forest structural parameters described in this chapter are mainly woody structural properties such as tree height, stem height, diameter at breast height (DBH), and forest stand density. 100 Optical remote sensing techniques have been applied in the retrieval of forest structural parameters. Tall vegetation species create larger shadows, but they also expose a larger portion of their vertical structure to the sensing systems when viewed from off-nadir directions. Consequently, optical sensors with off-nadir view angles can be used to study the vegetation biophysical structures (Qi et al. 1995). Li and Strahler (1992) built a geometric-optical model in which a tree canopy was a collection of individual geometric objects that cast shadows on a contrasting background: sunlit crown, shaded crown, sunlit background and shaded background. In this model, the three-dimensional structure of the canopy is the primary factor influencing the reflectance from the canopy (Strahler and J upp 1990). A good way of forest structure retrieval with optical remote sensing data is the inversion of the geometric-optical model (Woodcock et al. 1994; Shimabukuro and Huemmrich 1995). Scarth and Phinn (2000) defined forest structure as the horizontal and vertical distribution of components within a plant community. They applied TM images, DEM, and field measurements in an inverted Geometric-Optical model (Li and Strahler 1992) to estimate crown diameter, tree density, crown cover projection, and the “treeness” parameter. Recent developments in high spatial resolution remote sensing techniques leads to small scale characterization of canopy structure such as small forest gaps and individual crown characteristics. These structural properties are important in monitoring forest changes resulting from selective logging activities (Bongers 2001). 101 Microwave remote sensing techniques also have been applied to retrieve structural information in temperate forests. Previous studies have demonstrated that forest structural variation may have a substantial effect on L-band backscatter of forest stands with same biomass (Imhoff 1995). The spatial pattern such as inter-tree gaps and tree clumping in uneven forests may create significant differences in radar backscattering at finer resolution (Sun and Ranson 1998). The forest structural information retrieved from SAR backscattering is wavelength- dependent. Signals with shorter wavelength (e. g. X- and C-bands) mainly contain information in crown layers. More information about woody structures and biomass could be retrieved from longer wavelengths (e. g. L- and P-band) because the signals can penetrate the crown layer (Le Toan et a1. 1992). The backscattering is also polarization- dependent. The backscattering in each polarization contains different structure information. HH-polarized backscattering is dominated by the ground-canopy interaction mechanism while direct crown backscattering contributes more significantly to the V\/- and HV-polarized backscattering (McDonald and Ulaby 1993). Karam et al. (1995) also showed that the HV backscattering coefficient was due almost solely to the long branches at all ages. Therefore, the dynamic range observed in HV polarization reflected the physiological changes of long branches as the canopy age increased. The phase difference (the phase of the complex polarimetric coherence) in polarimetric SAR data, resulting from the crown attenuation and stalk-ground double bounce, contains much forest structural information (Ulaby et al. 1987). Phase difference could be 102 calculated in a polarimetric backscattering model (Ulaby et al. 1990). Hoekman and Quifiones (2002) proposed a new method of polarimetric coherence decomposition instead of power to relate polarimetric signal and forest structures. Structural information can also be described by image texture, the spatial variability of image tones that describes the relationship between elements of surface cover (Wulder et al. 1998). For images over forests, the variation in texture is related to changes in the spatial distribution of forest canopies. The medium to coarse resolution optical remote sensing imagery often has a smooth texture so that the information in texture variation is limited. However, the high resolution imagery such as IKONOS is able to reveal detailed variation of image texture. These data can be used to detect internal stand shade conditions as mutual shadowing that is more dominant in mature forest than secondary regrowth. SAR signal penetrates forest canopies much deeper than sunlight and therefore, contains more structural information than optical imagery. Luckman et al. (1997) showed that coefficient of variation (CV) of SAR backscattering intensity exhibited the least variation within each regrowth age, and was the best method to measure texture. Aside from the deficiency of speckle noise, the standard deviation of SAR intensity and SAR image texture is useful in forest structure retrieval (Manninen and Ulandder 2001). In this chapter, a microwave/optical synergistic radiative transfer model was developed and the backscattering from forest components was simulated while the leaf properties were measured from optical remotely sensed data. The forest structures were the input parameters of the model. The model was validated with J ERS-l SAR and VNIR data and 103 ground measurements. Then, with a nonlinear least square (LS) optimization method, the model was inverted and the forest structures estimated. Finally, an error analysis was performed to examine the accuracy of the model inversion. 5.2 Model Development The forest can be simplified as a three-layer vegetation composition: leaf canopy, branch+leaf canopy, and trunk atop of a rough soil surface (Figure 5-1). For different forest types, the components in each layer and the values of their biophysical parameters may vary. The saplings and seedlings and the grass/shrub underneath the forest are not considered in the simulation. The backscatter intensity is the result of additive contribution from the components in all layers: 0' = O-layerl + O-IayerZ + O-Iayer3 + 050i! (5-1) total In a 2nd-order solution of radiative transfer equations, the total backscatter consists of surface scattering from soil underneath, volume scattering from leaves, branches, trunks, interaction between forest components and soil surface, and 2nd-order non-coherent scattering: 0' : 050i] + Gleaf + abranch + 0' total trunk (5-2) + aleaf—soil + O-branclz-soil + O- + U trunk —soil noncoherenr The microwave canopy scattering model developed by Karam et al. (1995) had a detailed explanation of the scattering properties of the canopy-related components in , in Eq.5-2 is a very simple lst-order forests. However, in Karam’s model, the 0' solution of radiative transfer equations, assuming the soil as a continuous and slightly rough dielectric surface. It is thus insufficient to simulate radar backscatter under various 104 soil roughness and moisture conditions in different forest types, particularly in sparse, young, secondary forests in which the soil surface plays an important role in determining the total backscatter. Three modifications to Karam’s model were made in this study. The first was to introduce an integral equations model (IEM) bare surface scattering model (Fung et al. 1992) into the canopy scattering model. The second was to link the model to optical remote sensing variables. Since Karam’s model was originally developed for conifer forests, the probability distribution function (PDF) of each of the forest components was also modified to fit the characteristics of tropical forests in the study area. 5.2.1 Modified Karam-[EM model The first-order solutions of radiative transfer equations for backscattering coefficients (in power units) of the components in Eq.5-2 are described in Eq.5-3 to Eq.5-12. The second-order solutions of the non-coherent backscattering are much smaller than the first-order ones and, therefore, the equations are not listed in this study. 5.2.1.1 Soil surface scattering Soil surface backscattering in forests depends on its dielectric constant, surface roughness, and attenuation from the forest canopies above the soil surface. Surface roughness can be described with a root-mean-square (rms) of the surface roughness height and an auto-correlation function. The scattering coefficient of soil ground scattering (in power units) can be expressed as: 105 _ S 0'30” — tlprzpr3p0'pqr3qz'2quq (5-3) where p and q are polarizations (H or V), r, is the polarized attenuation factor from the ith layer (leaf canopy, branch+leaf, and trunk), and of”, is the pq element of the surface scattering matrix given in the [EM model (Fung et al. 1992). The IBM model extends its application to a wide range of soil surface conditions. It reduces to the small perturbation model when the surface is slightly rough and to the Kirchhoff model when surface roughness is large. Based on an approximate 2nd-order solution of a pair of integral equations for the tangential surface fields, the soil surface scattering is composed of single and multiple scattering when the surface rms slope is large: 5 _ single multiple _ 0m _ 0m +0120 (5 4) Since most natural terrain has a small rms slope, single scattering dominates over the multiple scattering in most situations (Shi et al. 1997). The single scattering coefficient (in power units) can be described as (Fung et al. 1992): 2 W" (‘ka ,0) n! n PC] sin le k2 —2/r252 00 Zn 0' g :78 2 ES 1 M (5’5) n=l where k is the wave number, k, = k cos 7] , k, = k sin 77 , 77 is incident angle, 3 is the rms height of the soil surface, I is the scattering intensity, and W" is the Fourier transfomi of the nth power of the surface auto-correlation function. 106 Multiple scattering is significant only when the surface rms slope is large. It dominates the cross-polarization backscattering. The multiple scattering coefficient (in power unit) can be described as (Fung et al. 1992): k252)n+m Umultiple _ k2 6—3/(2252 Z Z (2 pq _ 47:6 I I n: lm= l n’m' - IRCIfgqu (u,v)]W" (u - k,.,v) - W'" (kx + u,v)dudv ) [1+]?! 3: -21.,22 (kis +167: 22 n_1 m_ 1 n! my . [[le, (u,v)|2 + 17;, 01,”ng (—u,-—v)] . W’" (u — k,.,v)W" (u + kx,v)dudv (5-6) where f M and F p q are the coefficients of the Kirchhoff component and its complementary component in the electromagnetic field on soil surface. They are described in detail in Fung et al. (1992). 5.2.1.2 Leaf scattering In Eq.5-2, the leaf volume scattering coefficient (in power unit), 0,6,0] , is an additive contribution from all of the leaves in the first (JIM) and second (01:4) layers in forests. It is a lSt order solution of the radiative transfer equation (Karam et al. 1995): _ l 2 Glen] _ Clea] + aleaf _4flQIeafl . 1_Tlprlq +4 lea/2 . 1_T2Pr2(1 -T 2' _ ,,., k+k) ,,., k+k) ‘P' ( lp lq 86C]? ( 2p 2q seen (5-7) ‘I where k0. (i=1,2,j=p, q) is the extinction coefficient in the ith layer and jth polarization. Qi,”! and Qi,?” represent scattering from one leaf within layer 1 and layer2, 107 2 . . . . Q55] = nl -, where F 11:” IS the element of the scattermg amplitude matrix for a ' leaf, and n 1 is the number of leaves in this layer. The ensemble average <> is used in the equation to represent the statistical average over all leaves in this layer. 5.2.1.3 Branch scattering Similarly, the branch volume scattering in Eq.5—2, abram , is the 1St order solution of radiative transfer equation describing the scattering from branches in the second layer: l—z'zprzq _ branch Ubranch _ 472.qu . . z.1p2-1q (5-8) (k,p+k2q)sec77 branch _ branch 2 branch - . - - Here Q m _ n2 2, F M 1s the element of the scatterlng amplitude matrlx for a branch in layer2, and n2 is the number of branches in this layer. 5.2.1.4 Trunk scattering The description of trunk volume scattering in Eq.5-2, 0mm, , is similar with Eq.5-7 and Eq.5-8 except the attenuation factors: 1 — 2' 2' _ trunk 3P 3‘} Ummk _ 47Zqu I k k . 1'”)qu . zI2p’z-2q (5-9) ( 3p + 3g ) sec 77 Here Qtrunk = n3 ', F p?” 18 the element of the scattermg amplltude matrlx for a branch in layer3, and n 3 is the number of trunks in this layer. The volume backscattering of trunks suffer from the attenuation from both layer 1 (71) and layer2 (12). 108 5 .2. l .5 Leaf—soil interaction When calculating the interactions between scatterers in different layers, the soil surface is assumed to be a specular or a slightly rough surface to simplify the radiative transfer equations. In Eq.5-2, the backscattering coefficient caused by the interaction between leaves and soil surface, 0',eaf_30,, , is an additive contribution from the leaves in first (leaf canopy) and second (leaf+branch) layers. It can be described as: O-leaf—soil :. O-leafl-soil + 0160f 2—soil (5-10) where rlp—r k -4kzszcoszn . Qleafl . 14 . T 3qz’2q (7,,”ka = 4acosn-rlprzpr3p 2(Rpp +qu)-e M k lq- lp _4kzs2 c0327] . Qleafz . TZP — T24 pq ' 3a la k2q —k2p 0",,af2_m., = 47: cosn-z'lprzpr3p -(Rpp +qu)-e Here R pp and R W are the Fresnel reflect1V1ty of the 5011 surface at p or q polarlzatlon. 5 .2. 1 .6 Branch-soil interaction Similarly, abm,,c,,_soi, in Eq.5—2 can be described as: —4kzszcoszq branch TZP—Tzq 6 -Q -——-r,,r,q (5-11) Ubranch—soil = 47: COS 77 . Tlpz-ZpZ-iip . R Pq k pp 2q — 2p 5.2.1.7 Trunk-soil interaction Similarly, 0",,.unk_soi, in Eq.5-2 can be described as: _ 4 . R . —4k252coszn . trunk . 2.3P _ T30 5 12) Urrunk—soil — ”C0877 TlpTZpZ-3p . pp 6 qu .TZqZ-lq ( - k3,] — k3,, 109 5.2.2 Linkage to optical remotely sensed variables Optical remotely sensed data have been successfully applied to extract green forest biophysical attributes such as leaf area index (LAI), an indicator of green leaf biomass. The amount of green leaves controls the magnitude of attenuation from the leaf canopy in the canopy scattering model. Therefore, linking optical remote sensing variables to the model could provide quantitative information for the estimation of forest woody biomass. As described in Eq.5-3 te Eq.5-12, there is an attenuation factor in each layer of the forest (two factors in the second layer coming from the branches and leaves, respectively). Assuming the layer height is H, the attenuation factor in either forward or backward direction could be described as: T : e—k,H/cosr7 (5-13) I where t= p or q, the polarization status of the signal. The extinction coefficient (k,) is the total extinction cross section of all leaves statistically averaged over the orientation probability distribution: k = 4”NIm (5-14) I 0 where F a is the ce—pelarizatien component of the scattering matrix, or scattering amplitude tensor, of a single scatterer. N is the density (#/m3) of the scatterers in this layer. 110 In the canopy scattering model, the leaf is modeled as an ellipse. Let a and b be the half- length and half-width of an elliptic leaf respectively, the leaf density number N in Eq.5- 14 can be related to the leaf area index (LAI) by: LA] N: (5-15) 7rab(H1 +H2) It was assumed the same leaf density in first and second layers. In tropical forests, LAI can be estimated from green fractional cover (fc) that is derived from optical remotely sensed data as described in chapter 4: LA] = 0.217 + 0.058 - e(f‘-O'O33)2'5'578 0.4 < fc < 1.0 LAI = 0.85 - fc 0 < fc s 0.4 (5-16) MSA VI — MSA VI open MSA VI — MSA VI canopy open in a linear unmixing model as described in where fc = Chapter 3. The value of fc is in the range of [0,1]. MSA V1 is the Modified Soil Adjusted Vegetation Index and the detailed description could be found in Qi et al. (1994). 5.2.3 PDF functions of forest components In the canopy scattering model, the total backscattering in one layer is the statistical addition of the backscattering from all scatterers with a probability distribution function (PDF). As shown in Figure 5-2, the distribution of leaves in tropical forests is assumed to be approximately spherical, the branches are more plagiophile, and the trunks are almost vertical. The PDF equations of each forest component are listed below: 111 PDFW = Isin 204/2 [07:] 3 7t 57: PDF = 3 / — ——,— 5-17 branch ICOS al (4) [6 6 :1 ( ) PDmek = cos6 01K??? [0,3] where 0t is the inclination angle of scatterers. For elliptic leaves, it is the angle between vertical and the normal vector of the disc in clock-wise direction. Branches and trunks are simulated as cylinders and CI. is the angle between vertical and the axis of the cylinder in clock-wise direction. The denominators in Eq.5-17 are the normalization factors that are the integration of each function in the limited range of (X values. 5.3 Model Simulation and Validation By replacing the simple soil scattering model with the IBM model and linking some biophysical parameters to optical data, a microwave/optical synergistic canopy scattering model was developed to simulate scattering from tropical forests. The study area is Mae Chaem Watershed as described in Chapter 2. The remotely sensed data in microwave spectrum were JERS—l Synthetic Aperture Radar (SAR) data, and optical spectrum were JERS-l VNIR (visible + near infrared) data. Only backscatter in L-band and HH polarization with incidence angle of 35° was modeled in order to match the JERS-l SAR configuration. 5.3.1 Remotely sensed data The study area covers two scenes of JERS-l SAR data (path/row): 132/269 and 132/270. Two scenes of J ERS-l VNIR data in the same orbit were also acquired. The JERS-l SAR 112 is an active sensor that transmits L-band microwave signals and detects the characteristics on the Earth surface with the backscattered signals. JERS-l VNIR is a high-resolution radiometer that observes targets by receiving solar radiation reflected from the earth surface in three visible and near-infrared (NIR) bands. The system parameters of these sensors are listed in Table 5-1. The SAR data was acquired 8 days earlier than the VNIR data. The surface change during this period was thus neglected. The two scenes of SAR or VNIR data are in the same orbit and therefore, it is easy to mosaic them into a larger scene (Figure 5-3). Beth SAR and VINR mosaic image covered most of the study area with gaps on the east and west edges. With the information in the metadata, both SAR and VNIR images were geerefereced using UTM projection (zone 47) with spheroid and datum WGS84. There was no cloud cover in the VNIR images. The atmospheric effect in VNIR images was corrected with a simple dark—object- subtraction (DOS) technique (Chavez 1988). SAR and VNIR images were geometrically corrected by setting the reference image as the ETM+ image acquired on February 2, 2000, which has been geometrically corrected with ground control points (GCP) collected during field trips. From a nadir view, the VNIR images did not have much geometrical distortion. The total error was 6.14m in a second-order polynomial geometric model. The JERS-l SAR is a side-looking sensor with a shallow incident angle. As a result, its imagery is severely distorted in the mountainous area. 113 SAR data was processed with a 3x3 Lee filter to reduce the speckle noise and increase the signal/noise ratio with increased number of looks. During the geometric correction, considering the large area of the coverage and the severe topographic variation, the study area was divided into four parts with a narrow overlay zone. Each part of the data was geometrically corrected by setting the ETM+ image as the reference. In a 3rd-order polynomial geometric model, the total errors are 51.68, 25.23, 42.13, and 48.78 meters for the lower right, upper left, upper right, and lower left parts, respectively. Since the look direction of JERS-l SAR was from west to east, the errors in west-east direction were much higher than north-south direction. The four parts were then mosaiced again (F ig.5-4b). Even splitting the total area into 4 sub-areas, it is still difficult to correct the geometric distortion in the watershed. In the areas with high elevation and slopes (Fig.5- 4a), the geometric correction has very high error. To constrain the total error in one pixel, the pixel size of geometrically corrected SAR data was resampled to 60 by 60m with a nearest neighbor method. SAR backscattering was also affected by the topographic variation in mountainous areas. There are several types of topographic distortions in the SAR image: foreshortening, layover, and radar shadow (Hendenson and Lewis, 1998). F ereshortening occurs when the ground range difference between two ground positions is reduced to the slant range difference, which is shorter because of the side-looking characteristics of SAR and the topography on the ground. When the shortening becomes greater, the position on the top of the mountain will be closer to the antenna than the position on the bottom, then 114 layover occurs. In this case, the radar signal cannot access the opposite slope, and radar shadow occurs. The foreshortening effect can be corrected with DEM data. During topographic correction, the backscattering amplitude, DN, of the SAR image can be corrected to a reference surface on which the local incidence angle of each pixel is 0 (Van Zyl et al. 1993) sin 77 cos 90 DN = DN* (5-18) topoc0r sin 770 where 770 is SAR incidence angle, 77 is local incidence angle, and «9,, is azimuth slope. Setting 6 as local slope angle and Ago as the relative azimuth angle between local aspect and SAR azimuth, local incidence angle?) is calculated: c0377 = cochos 770 + sinI9sin 770 cos Ago (5-19) The azimuth slope 60 can be solved from tri-angle relationships: tha = tgbl sin Ago (5-20) In the areas with layover and radar shadow, the SAR data were permanently lost and cannot be corrected. The conditions when layover and radar shadow occur are: Layover: 77 < 900 and 6 < 7; Radar shadow: 77 > 900 Figure 5-5 is the local incidence angle (a) and the topographically corrected image (b). The areas with layovers and shadows were shown as white color in the local incidence angle image. The backscattering amplitude in these areas cannot be used in the following 115 analysis. In high mountainous areas with large local incidence angles, although the layovers and shadows did not occur, the foreshortening effect was so severe that the correction accuracy was very poor. For this reason, the application of JERS-l SAR data in mountainous area is limited. 5.3.2 Model simulation The model was simulated with a set of parameters listed in Table 5-2. The range of each parameter was determined with the experience gained during field trips. The mean of the parameter was selected approximately based on the frequency of occurrences. To examine the contribution of these biophysical parameters on the backscattering, only the values of one parameter was changed during each simulation, while other parameters were represented by their mean values. With the configuration in Table 5-2, in L-band HH polarization, branch volume scattering, trunk-surface double bounce, and leaf volume scattering are the primary components in total backscattering (Figure 5-6). 5.3.2.1 Contribution of leaves In the microwave/optical synergistic forest scattering model, the leaves are simulated as elliptic discs of which a], a2 are the radii of long and short axis, and r=al/a2 is the elliptic ratio of the leaf. Figure 5-6a shows the variation of backscattering in different leaf size and elliptic ratio. When both the leaf size and elliptic ratio are small (a]=0.02m and r=l .2), the backscattering coefficient is very low and then increases rapidly with higher elliptic ratio. When the leaf size is larger, the backscattering decreases slowly with elliptic ratio. With much larger leaves (a1=0.07), there is almost no difference in the 116 backscattering with the change of elliptic ratio. The backscattering difference in the range of leaf sizes could be more than 2dB in small elliptic ratio (r=1.2) while only 0.3dB in large ratio (r=1.5). Therefore, a larger leaf size (a1=0.05) and elliptic ratio (r=1.5) is used in the following simulations. The contribution of different components in the forest model is shown in Figure 5-6b with the change of LAI, which is directly related to leaf size and canopy height (Eq.5-15). When LAI values increase, both the leaf volume scattering and leaf-soil interactions increase rapidly then slow down to reach the saturation point. The contributions of other components decrease due to the attenuation from the leaves. As described in Eq.5-13 and 5-14, the attenuation loss in dB unit is more or less linearly related to the amount of leaves. With the configuration listed in Table 5-2, the increased backscattering from leaves are compensated with the backscattering loss from other components, and therefore, the total backscattering increase is only ldB in Figure 5-6b. 5.3.2.2 Contribution of branches The branch-related backscattering varies with both branch size and density. The branch length is assumed to be same as the height of layer 2. As shown in Figure 5-6c, in L-band HH polarization, the modeled branch volume scattering and branch-soil interaction are very sensitive to branch radius. The branch volume scattering has the highest value at branch radius = 0.018m and lowest value at radius = 0.028m, and the difference is as high as 8dB. The branch-soil interaction has an obvious low value at radius = 0.033m, and the 117 difference between highest and lowest scattering is around 3dB. The backscattering of other components changes with a much smaller scale. With the increase of branch density (#/m2), both the branch volume scattering and branch-soil interaction have higher values, and the increment is 5dB and 3dB, respectively (Figure 5-6d). The branch-soil interaction is more quickly to reach saturation. The scattering of other components decreases almost linearly because of the attenuation in branch layer. 5.3.2.3 Contribution of trunks The trunk-related backscattering varies with both trunk size and density. Trunk size is determined by DBH and trunk height, and the trunk density is same as the stand density in the forest. In contrast with the contribution of leaves and branches, the trunk-soil double bounce is much stronger than trunk volume scattering. The trunk-soil interaction and trunk volume scattering are sensitive to the variation of DBH (Figure 5-6e). The difference of the highest and lowest values in trunk-soil interaction is as high as 9dB. The difference in trunk volume scattering is only 3dB, but the curve is more complicated. The leaf and branch volume scattering is not affected by the variation of DBH. The scattering in other components decreases almost linearly. In accordance with the curve of trunk-soil interaction, the total scattering increases slowly, then decreases after DBH > 0.16m. The change in total scattering is 2.6dB. 118 As shown in Figure 5-6f, when stand density (# of trees / m2) increases, the trunk-soil interaction increases, then decreases after stand density > 0.08. The possible reason is that, with denser trees, there is less space for double bounce between trunks and soil surface. The decrease is around 3dB in the range of stand density in Figure 5-6f. On the other hand, the trunk volume backscattering increases up to 5dB. It is obvious that there is more chance of trunk volume backscattering at higher stand density. The leaf and branch volume scattering is not affected by the variation of stand density. The scattering in other components decreases ahnost linearly. With the compensation of increased trunk volume scattering, the total backscattering decreases only 1dB. It was found in ground-measured data that trunk height was almost linearly correlated with tree height (R2=0.81). Here the backscattering is simulated with the variation of tree height, which affects the heights of three layers (leaf canopy, leaf+branch, and trunk). These variations in turn result in different contribution from leaves and branches. In Figure 5-6g, the scattering from all components (except leaf volume scattering) is tremendous. When the tree height increases from 6m to 24m, the branch volume scattering, the component that has highest contribution to total scattering, increases up to 6dB. The trunk volume scattering increases slowly, then drops 7dB. The trunk-soil double bounce drops more than 40dB. The soil surface backscattering, leaf-soil and branch-soil interaction even reaches infinity (-80 dB in the model) when the tree is higher than 18m. With the compensation of decreased scattering from these components, the total backscattering increases 2dB. 119 5.3.3 Model validation Field measurements were made in January 20-27, 2002 during the second field trip, one week earlier than the JERS-l SAR data acquisition. Among the 32 study sites visited, there were 9 dry dipterocarp forests, 9 mixed deciduous, 5 pine transition, 6 dry evergreen, and 3 moist evergreen. With a point-quadrant plot method, the biophysical attributes at each site were measured: tree height, trunk height, DBH, stand density. The forest fractional cover and LAI values were estimated with JERS-l VNIR imagery. The microwave/optical synergistic canopy scattering model is validated with these ground measurements and JERS-l SAR data. In addition to the ground-measured biophysical attributes, other inputs of the model are listed in Table 5-3. The values of some parameters be different in each forest type. The leaf and branch radii and soil conditions are selected based on the experiences in fieldwork. The dielectric constants (8) of forest components are calculated based on vegetation moisture and local temperature with the model described in Ulaby et al. (1990). Only backscatter in L-band with incidence angle of 35° and incidence azimuth angle of 278° is modeled to be consistent with JERS SAR system. As shown in Figure 5-7, there are logarithmic relationships between modeled backscattering (LHH) and ground-measured biophysical parameters. The modeled backscattering increases rapidly and then reaches the saturation point at LAI = 4 (Figure 5—7a). The increments of modeled backscattering with tree height (Figure 5-7b), trunk height (Figure 5-7c), and DBH (Figure 5—7d) are also observable, but the residuals are 120 higher because of the attenuation from leaves. The relationship between modeled backscattering and stand density is very weak (Figure 5-7e). One possible reason is that stand density is not a good indicator of forest growth. A young secondary forest could have similar stand density as natural undisturbed forest. In late succession stage, the stand density of natural forest declines, a similar phenomenon as the degraded forests disturbed by human and natural activities (Aber and Melillo 2001). The modeled backscattering coefficients are compared with J ERS-l SAR measurements. In Figure 5-8a, all points were scattered along the 1:1 line, indicating that the model fits well with SAR data. The root mean square error (RMSE) of the model simulation can be calculated as: M 0 0 2 2(05/11? - O-mod e!) RMSE = M M (5-21) where M is the number of study sites that were used for validation. The RMSE in this study was 0.94dB in comparison with the JERS-l SAR observed and modeled backscattering coefficients at the 32 study sites. As shown in the residual plot in Figure 5- 8b, most of the residuals scatter between ildB. The only site with residual higher than 2dB is teak plantation in the study area, whose regular pattern results in very high backscattering in JERS-l observation. Figure 5-8 indicates that the synergistic canopy scattering model could be used to predict the backscattering in the study area at an accuracy of around 1dB. 121 5.4 Model Inversion and Forest Parameters Estimation 5.4.1 Model inversion In the microwave/optical forest scattering model, the backscattering coefficient is determined by several forest structural parameters: leaf size, leaf density (#fha), dielectric constants 8W , (9th , 8mm, , heights of each layer, radii of branches and trunks, and density of branches and trunks (#/ha). 0 _ . . atom] — f(leaf .. Slze’ leaf __ denSlty, gleaf. Sbranch.8rrunk. Hlaycrl, HlayerZ, Hlayer3 ’ branch _ radius, branch _ density, DBH, 5 tan d _ density) (5-22) The forest structures are unique in each forest type. In accordance with the forest type map created from optical remotely sensed data, only three major forest types are considered in this chapter: dry dipterocarps, mixed deciduous, and tropical evergreen forests. Table 5-4 lists a series of parameters that are applied in the model for each forest type. As described in Chapter 2, some of the forest structural parameters are highly correlated to each other. Figure 5-9 is the scatterplots of ground-measured stem height vs. tree height and DBH vs. tree height. The regression equations are: stem_H = 0.559*tree_H —l.328 (5-23) DBH = 1.332 *tree__H + 4.29 Here the units of stem_H and tree_H are meters while the DBH is in centimeters. The leaf density can be calculated with Eq.5-15 and 5-16 from optical remotely sensed data. With the relationships in Table 5-4 and Eq.5-23, the model only needs two input parameters: tree height and forest stand density. Then Eq.5—22 becomes: 0'0 = f (tree_ H, s tan d _ density) (5-24) mod e! 122 For each pixel in the JERS-l SAR data, a least-square-error optimization technique was applied to estimate these two forest structural parameters with model inversion. The standard criteria listed below: - 0 0 2 mln(0mod el - O-SAR) The initial values of tree_H and stand_density for each forest type were given at the first iteration. At each iteration, a new 0210“, was generated and compared with GEAR (J ERS- l) at each pixel. The model stopped when the derivative of 0(3),”, between two iterations was less than 10'6 or the iteration numbers exceeded 104. Then the tree_H and stand_density at that iteration were outputted. 5.4.2 Forest structural parameters by model inversion Figure 5—10 is the modeled tree height distribution in Mae Chaem Watershed. The non- forest areas as shown in the forest type map (Figure 2-4) are masked out in Figure 5-10 and the following figures. Most of the trees in the watershed are higher than 10 meters. Most of the dry dipterocarps are lower than 15 meters. Only some isolated dry dipterocarps are around 20 meters. The mixed deciduous forests have a wider range between 15 to 30 meters and the distribution is more heterogeneous than dry dipterocarp forests. There are no trees higher than 30 meters in dry dipterocarps and mixed deciduous forests. The tropical evergreen forests have a very wide range changing from 15 meters to higher than 30 meters. There are also some isolated areas with tree heights around 12 meters. 123 Figure 5-11 is the modeled forest stand density distribution in the watershed. In dry dipterocarps and mixed deciduous forests, there is a wide variety of stand density ranging from 100 to 600 trees/ha and higher. The stand density of dry dipterocarps is much higher than mixed deciduous. In accordance with the ground-measured parameters, the stand density is negatively related to tree height. When the tree is higher than 15 meter, the stand density in this area is around 100-400 trees/ha. When the tree is lower than 15 meter, the stand density could reach 600 trees/ha or higher. Opposite with the tree height map, there is a trend of decreasing stand density from east to west in mixed deciduous forests. The model does not work for tropical evergreen forests. The modeled stand density distribution saturates at around 500 trees/ha. 5.4.3 Uncertainty analysis The error 8 of the model inversion is defined as the absolute difference between modeled and JERS-l SAR observed backscattering coefficients. It is in the unit of dB: 0 — 0.2/1R (5-24) ‘9 : Umodel In Figure 5-12, most of the areas in dry dipterocarp forests in the watershed have errors less than 1dB, indicating that the forest structural parameters estimation with model inversion works well in these forests. In the mixed deciduous forests where the topographic relief is high, the error could reach 2-3 dB. In some areas on the tops of mountains, mostly in tropical evergreen forests where the elevation is very high and slope is very steep, the error is even higher than 4dB. In these areas, the tree height and stand density are highly overestimated or underestimated. As shown in Figure 5-5b, the 124 topographic correction of the SAR image did not work well in these areas, which introduces high errors from topographic effects. The modeled tree height and stand density were also compared with ground-measured values at the study sites. The average of the model inversion error at all study sites was only 0.44dB although the individual values range from 0 to 3.26 dB. In Figure 5-13a, the modeled and measured tree height values at most of the study sites scattered along the 1:1 line. There was obvious overestimation in young forests in which the trees were shorter. These forests were mostly dry dipterocarps in which the measured trees were between 5- 10 meters whereas the modeled ones were 10-15 meters. For forests with higher trees, the modeled and the measured values fit well. There is one outlier in Figure 5-13a with low value of measured tree height (13.18 meter) but high value of modeled tree height (28.0 meter). It is a degraded and recovered tropical dry evergreen forest with high heterogeneity at the study site (82-14). The ground measurements could be underestimated using point-quadrant plot method as discussed in Chapter 2. On the other hand, the canopy scattering model did not account the scattering contribution from the interaction between young (short) and old (tall) trees at this site. As a result, a higher tree height was introduced to match the JERS-l SAR observed scattering coefficients. The error of model inversion at this study site is only 0.19dB. 125 The measured and modeled stand density at the study sites also scattered along the 1:1 line (Figure 5-13b). The modeled stand density at most of the tropical evergreen sites, including the outlier (S2-14) in Figure 5-12, saturate at about 500 trees/ha. For other forests (mostly dry dipterocarps and mixed deciduous), the modeled and measured stand density fit well. For each forest type, the average and standard deviation of modeled tree height, stand density and the error of model inversion at all study sites were listed in Table 5-5. The modeled tree height increased from dry dipterocarps, mixed deciduous, pine transition, to tropical evergreen forests. The pine transition had the highest standard deviation. However, in the forest type map, the pine transition, dry evergreen, and moist evergreen forests were combined as tropical evergreen forests. In this way, the standard deviation is only 2.93. In accordance with ground measurements, there is a negative relationship between tree height and stand density. Young secondary dry dipterocarps have higher density, while the tropical evergreen forests with less human disturbance have lower density. The standard deviation of moist evergreen forests is very low (0.58) because of the saturation of model inversion in these forests. Since most study sites were chosen at relatively flat areas, the topographic effect at these sites was minimal and the average errors of model inversion at all forest types were less than 1dB. The pine transition had the highest standard deviation that leaded to a model inversion error of around MB in tropical evergreen forests. The model inversion in dry dipterocarps and mixed deciduous forests worked very well. 126 In Figure 5-14, the average values of modeled tree height (Figure 5—14a) and stand density (Figure 5-l4b) were compared with the ground-measured values in each forest types. The modeled tree heights were generally overestimated. The overestimation was more obvious in dry dipterocarps and moist evergreen forests, and less obvious in mixed deciduous, pine transition, and dry evergreen forests. The total root-mean-square-error (RMSE) in tree height estimation was 4.6 meter in all study sites, and 3.8 meter when S2- 14 was removed. The modeled stand densities in dry dipterocarps and pine transition fit well with ground-measured values. But they were overestimated in mixed deciduous and dry evergreen forests, and were underestimated in moist evergreen forests because of the saturation in model inversion. The total RMSE in stand density estimation was 300 #/ha in all study sites, and 299 trees/ha when 82-14 was removed. 5.5 Conclusions and Discussion In this chapter, a microwave/optical synergistic radiative transfer model was built to simulate the radar scattering from forests in the study area. The backscattering coefficients of the forest components, such as branches and trunks, were modeled. The leaf scattering and its attenuation to the woody components were quantified in the model with leaf area index retrieved from optical remotely sensed data using the method in Chapter 4. The root-mean-square error (RMSE) of the model was 0.94 dB when compared with JERS—l SAR backscattering coefficients. Two forest structural parameters, tree height and stand density, were estimated by model inversion with least-square optimization techniques and JERS-l SAR and VNIR data. 127 The error of model inversion was less than 1 dB in most of the areas in the watershed. However, in the areas with high relief and steep slopes, the model inversion error could be higher than 4 dB, which introduced high error in the estimation of forest structural parameters. When all study sites were considered, the total root-mean-square error of tree height estimation was 4.6 meter. The total root-mean-square error of stand density estimation was 300 trees/ha. The average model inversion error was 0.44 dB although the individual values ranged from 0 to 3.26 dB. The model worked well for the estimation of structural parameters in dry dipterocarps and mixed deciduous forests. Most of the dry dipterocarps were lower than 15 meters. The mixed deciduous forests had a wider range between 15 to 30 meters and the distribution was more heterogeneous. The stand density in these forests varied from 100 to 600 trees/ha or higher. In accordance with the ground measurements, the tree height was negatively related to the stand density in these forests. The error of model inversion was less than ldB in most of these forest areas. The modeled tree height in tropical evergreen forests ranged from 15 to 30 meters and higher. The model inversion to estimate stand density did not work well in these forests. Its distribution saturated at approximately 500 trees/ha. In mountainous areas, the estimation of forest structures with model inversion was highly affected by the quality of SAR imagery and the accuracy of topographic correction. The evergreen forests, especially moist evergreen forests, are distributed on the top of mountains with high relief. It was difficult to correct the topographical effects to JERS-l 128 SAR data in these areas. The error of model inversion in tropical evergreen forests was higher than that in other forests. In some areas when the relief was very high, the topographic effect cannot be corrected. The errors of model inversion in these areas were higher than 4dB, which introduced high uncertainty in the estimation of forest structural parameters in these areas. A more accurate DEM data and a better correction algorithm technique for the topographic correction to SAR imagery will be studied in the future. The dense leaf cover in tropical forests is a big obstacle in SAR remote sensing. It highly attenuates the scattering from woody components like trunks, branches, and their interaction with soil surface. The model developed in this study took the advantage of leaf properties quantified with optical remotely sensed data. As a result, the modeled backscattering was more sensitive to woody parameters in tropical forests. Therefore, this model can also be used for the woody biomass estimation in tropical forests which is discussed in next chapter. 129 5.6 References Aber, J. D., and A. Melillo (2001). Terrestrial Ecosystem. HP Harcourt Academic Press. Barkman, J. J. (1979). The investigation of vegetation structure and texture. 123-160. In: Werger, M. J. A. (ed.), The study of vegetation. Dr. W. Junk Publishers, Hague. Bongers, F. (2001). Methods to assess tropical rain forest canopy structure: an overview. Plant Ecology: 153, 263-277. Chaves, P. S. Jr. (1988). An improved dark-object subtraction technique for atmospheric scattering correction of multispectral data. Remote Sensing of Environment: 24, 459-47 9. Fung, A.K., Lee, 2., and Chen, K. S. (1992), Backscatter from a randomly rough dielectric surface. IEEE Transactions on Geoscience and Remote Sensing: 30 (2): 356-369. Hoekman, D. H. and Quifiones, M. J. (2002). Biophysical forest type characterization in the Colombian Amazon by airborne polarimetric SAR. IEEE Trans Transactions on geoscience and Remote Sensing: (40)6, 1288-1300. Imhoff, M. L. (1995). A theoretical analysis of the effect of forest structure on Synthetic Aperture Radar backscatter and the remote sensing of biomass. IEEE Transactions on Geoscience and Remote Sensing: 33(2), 341-352. Karam, M. A., Amar, F., Fung, A. K., Mougin, E., Lopes, A., Le Vine, D. M. and Beaudoin, A. (1995). A microwave polarimetric scattering model for forest canopies based on vector radiative transfer theory. Remote Sensing of Environment: 53, 16-30. Le Toan, T., Beaudoin, A. and Guyon, D. (1992). Relating forest biomass to SAR data. IEEE Transactions on Geoscience and Remote Sensing: 30, 403-411. Li, X., and Strahler, A. H. (1992). Geometrical-optical bidirectional reflectance modeling of the discrete-crown vegetation canopy: effect of crown shape and mutual shadowing. IEEE Transactions on Geoscience and Remote Sensing: GE-30, 27 6- 292. Luckman A. J. (1997). The effects of topography on mechanisms of radar backscatter from a coniferous forest and upland pasture. IEEE Transactions on Geoscience and Remote Sensing: 36(5), 1830-1834. Manninen, A. T. and Ulander, L. M. H. (2001). Forestry parameter retrieval fiom texture in CARABAS VHF-band SAR images. IEEE Transactions on Geoscience and Remote Sensifiz: 39(12), 2622-2633. 130 McDonald, KC. and Ulaby, F. T. (1993). Radiative transfer modeling of discontinuous tree canopies at microwave frequencies. International Journal of Remote Sensing: 14(11), 2097-2128. Qi, J ., Cabot, F ., Moran, M. S. and Dedieu, G. (1995). Biophysical parameter estimations using multidirectional spectral measurements. Remote Sensing of Environment: 54, 71-83. Search, P. and Phinn, S. (2000). Determining forest structural attributes using an inverted Geometric-Optical model in mixed Eucalypt forests, southeast Queensland, Australia. Remote Sensing and Environment: 71, 141—157. Shi, J. C., Wang, J., Hsu, A. Y., O’Neil, P.E., Engman, E. T. (1997), Estimation of bare surface soil moisture and roughness parameters using L-band SAR image data. IEEE Transactions on Geoscience and Remote Sensing: 35:1254-1265. Shimabukuro, H. F. and Huemmrich, K. (1995). Remote sensing of forest biophysical structure using mixture decomposition and geometric reflectance models. Ecological Application: 5(4), 993-1013. Strahler, A. H. and Jupp, D. L. V. (1990). Modeling bidirectional reflectance of forests and woodlands using Boolean models and geometric optics. Remote Sens'mg and Environment: 34, 153-178. Sun, G. and Ranson, K. J. (1998). Radar modeling of forest spatial patterns. lntemational Journal of Remote Sensing: 19 (9), 1769-1791. Ulaby, F. T., Held, D., Dobson, M. C., McDonald, K. C. and Senior, T. B. A. (1987). Relating polarization phase difference of SAR signals to scene properties. IEEE Trans Transactions on geoscience and Remote Sensing: 25 , 83-92. Ulaby, F .T., Sarabandi, K., Mcdonald, K., Whitt, M. and Dobson, M. C. (1990). Michigan microwave canopy scattering model. International Journal of Remote Sensing: 11(2), 1223—1253. Van Zyl, J. J ., B. D. Chapman, B. D., P. Dubois, and J. Shi, (1993). The effect of topography on SAR calibration. IEEE Transactions on Geoscience and Remote Sensing: 31 (5), 1036-1043. Woodcock, C. E., Collins, J. B., Gopal, S., Jakabhazy, V. D., Li, X., Macomber, S., Ryherd, S., Harward, V. J., Levitan, J., Wu, Y. and Warbington, R. (1994). Mapping forest vegetation using Landsat TM imagery and a canopy reflectance model. Remote Sensing of Environment: 50, 240-254. Wulder, M. A., LeDrew, E. F ., Franklin, S. E. and Lavigne, M. (1998). Aerial image texture information in the estimation of northern deciduous and mixed wood forest leaf area index (LAI). Remote Sensing of Environment: 64, 65-76. 131 Table 5-1 System parameters of J ERS-l SAR and NVIR sensors. J ERS-l satellite Altitude: Orbit: Orbital direction: Descending ~5 70km 97.67°, sun-synchronous SAR VNIR Spectral frequency L-band (1.275 GHz) Band]: 0.52 — 0.60 pm (green) Band2: 0.63 — 0.69 pm (red) Band3: 0.76 — 0.86 pm (NIR) View angle Incident angle (right-sided): IFOV: Near-range; 35° 21.3 (nadir viewing) Far-range: 42° Polarization HI-I / Pixel size 12.5m 18m Swath width 75km 75km # of Looks 3 / Acquisition date 02/27/1998 03/ 06/ l 998 Table 5-2 Input parameters for model simulation. parameter range mean Leaf size (m) 1/2 long axis 0.02 ~ 0.07 0.05 elliptic ratio 1.2 ~ 1.5 1.5 Branch size (m) radius 0.016 ~ 0.047 0.031 Trunk size (111) radius 0.06 ~ 0.22 0.1 LAI 0.2 ~ 8.0 1.0 Branch density (#/m2) 0.5 ~ 3.0 1.5 Tree density (#/mT) 0.03 ~ 0.18 0.1 Tree height (m) 6 ~ 24 10 132 Table 5-3 Input parameters for model validation (in addition to ground measurements). dry mixed pine dry moist dipterocarps deciduous transition evergreen evergreen Leaf radii (0.07,0.05) (0.05,0.033) (0.03,0.02) (m) (31,32) Branch 0.022 0.029 0.031 0.033 radius (m) 8133f (l6.94,-5.7) (22.29,-7.26) (28.23,-8.95) (35.23,- 10.78) ebmnch (19.7,-6.53) (34.45,- 10.49) gmmk (13.57,-4.66) (26.53,- 8.41) Sson (5.3,-0.65) (10.23,- 1.95) Soil Root-mean—square height (m): 0.015 roughness Correlation length (m): 0.15 Table 5-4 A set of forest structural parameters for each forest e in the model. dry dipterocarps mixed deciduous tropical evergreen Leaf3-D size (7, 5, 0.5) (5, 3.3, 0.5) (4, 3, 0.7) (cm) 812a! (l6.94,-5.7) (22.29,-7.26) (28.23,-8.95) gm” (5.3,-0.65) (5.3,-0.65) (10.23,-1.95) 3mm (19.7,-6.53) (19.7,-6.53) (26.53,-8.41) grrunk (l3.57,-4.66) (13.57,-4.66) (23.26,-7.53) Height of layer 1 (tree_H- (tree_H-stem_I—I)2*/ 3 (tree_H- stem_H)*2/ 3 stem_H)/ 3 Height of layer 2 (tree_H- (tree_H-stem_H)* 1/ 3 (tree_H- stem_H)* 1/ 3 stem_H)*2/ 3 Height of layer3 Stem_H Stem_H Stem_H Branch radius DBH/4 DBH/4 DBH/4 Branch density Stand_density* 8 Stand_density* l 6 Stand_density*32 133 Table 5-5 Average and standard deviation of modeled tree height, stand density, and the error of model inversion for each forest types. tree H density error (m) stddev (#/ha) stddev (dB) stdev Dry dipt 14.34 2.35 701.11 328.05 0.13 0.40 Mixed dec 14.65 0.58 677.33 313.01 0.18 0.46 Pine trans 16.73 5.93 747.80 342.98 0.76 1.41 Dry ever 15.65 2.45 639.80 313.16 0.74 0.66 Moist ever 25.99 0.41 499.33 0.58 0.50 0.86 134 Figure 5-2 Figure 5-1 Geometry of a forest canopy. 1.4 _ 2222222 trunk leaf 2 2 - - — branch 1.2 2 1 a l‘\. . lfi \ / O 8 / x, ,/ \. B ' A ‘,‘ / \' ’I \. D. . [I \‘ 'I \‘ 0.6 — I \ I 1 ‘~ I .‘ .1 \ I \. , 0'4 I 2/ I I \ ‘2’ I I. ’1' 0.2 2 .’ 'I I ,' I l -- x I. 0 I J, I I I I I I ‘ I 0 20 40 60 80 100 120 140 160 180 inclination angle PDF functions of leaves, branches, and trunks applied in the synergistic canopy scatter model. 135 Figure 5-3 JERS-l SAR (a) and VNIR data (NIR+Red+Green) (b) in the study area. 136 z. . 1-1 ‘ xiv-.314 ' ,I Figure 5-4 DEM data (a) and geometrically corrected JERS-l SAR image (b). Figure 5-5 Local incidence angle image (a) and the topographically corrected image (b). The layovers and shadows are the white areas in (a). 137 -8.5 I I l r I A '9 ‘ :-: : ‘ A m F ‘ :':-~. .3 ----------------------- -----‘--$.'.;_;~--.- e .5? 0'1 -9 5 W "In- '0 2 - — - , - g -10 - _/ - I g ./ ——-a1=0.07 '/ ........ a1=0.06 40.5 r- ,/ -—-- a1=0.05 ./ —-a1=0.04 ,/ ----- a1=0.03 / ._-— = -11 2' I J J 1 31 0.02 modeled sing (dB) modeled slgmo (dB) modeled sigmo (dB) 1.2 1.25 1.3 1.35 1.4 1.45 1 leaf elliptic ratio (a1/a2) .5 '5 l I l I I l -10 _\M - ’ . ' ‘ ‘ ' ‘ - - - , ______ c 45 _ ,- .......... _ ( ) -20-fi_____.______~::"__“: .. -25 — — —total " '”’ ‘— " ‘— '— ~\ —\ _\ -— soll - - leaf '30 ‘ 232:“; -_v -_-_____,‘.: --------- ‘ ----- leaf-soil “~-"’ "“‘"--—- ----- branch -35 --_ ______________________ .. ---- branch-soil -------- —-- --—~ trunk -------- trunk-soil _40 l l l l l l 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.05 branch radius (m) '5 I T u— r 1 r -10 -fl— _ -15 _ —I (e) -20 ________._, _________ _ '25 "' —.§_ -\ ‘\ —\ ‘ ----------- total -30 _ _ _,‘- f \ -\_ a» ‘___ ._ — -— soil /“‘~\:_~/:"_" :~ * \ -2 — — :93; 'l - _ ----- ., ‘\-“ _ ----- ea-sor 35 “ ------- y r - c ----- brancE I "‘-~-_-,_ ---- branc -soi -40 — ------- ‘ --—— trunk ~ ------ trunk-soil -45 M 1 1 1 1 1 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 DBH/2 (m) 0 I l r '10 ‘ _____f_. ----------- "I (g) -20 _ r -------------------- a \ ‘\ '30 r “::-:;=.:_\_ ,,,,,, a ‘ “\?*->: “"~ ~- -- '40 ' ‘~.“‘\\ “ 1 ——total _ ‘\ _ -- sell 502 .\\ __leaf ‘. ‘ ----- leaf-soil '60 5 ‘. *\ 4 ----- branch '~ ---- branch-soil -70 — z, \ - ----trunk 2 --------- trunk-soil -80 I I ~ 1 I 5 10 15 20 25 tree height (m) 138 modeled slng (dB) modelded sigmo (dB) modeled slgmo (dB) 1 1 I I I T I -10 _ (b) -15 _ , ’ ’ I , v - ---- ‘- .{_ / -20 - / """" d / -25 J — __~ -_ ‘‘‘‘‘‘ total -30 P - --------- _ T. :——-— v; 2‘; 7:: '* SOII v“.\'“~r~;‘; "‘- _ . — _ 'eaf ‘35 _ x . runawa‘o“ ~~....- ----- leaf-soil ----- branch , .—.- branch-soil 40 ~, ‘1 ---- trunk -45 1 1 1 1 1 1 1 - ...... UUHK’SOII J 0 2 4 6 8 LAI -5 I I I I -10 W, ---------- T -15 r — _m_ ~~--:‘:ii:fV -25 fl‘ .\ ‘ n —total \ _\ _\ ‘ -— soil 5 -\ ,\ __§ — — leaf -30 - ~‘ _--_2__ _______________________ f ----- leaf-SOil I"’--"¢_:::_-—- -‘\ ~~- “- ----- branCh -35 I __ _______ __, .......... ___ -—-- branch-S< _______ ~__~__ --—-trunk .40 1 1 1 1 ------- trunk's‘)", 0.5 1 1.5 2 2.5 3 branch density (#lm2) -5 1 1 1 1 r 1 —10 - -15 — d ,\ ‘\ -25 r ‘\ ‘\ - _ —I0t?' ‘ \ ‘\ -— sell \ \ ‘ ‘ ‘\ ‘\ _ - laf ‘30 «Nun- \ ..‘,~ ,_ __-___-.:-7=-=';: """ leaf-soil ’I-‘Lh;-:':-' ‘ T ‘ - ..... branCh -35 ~ I, “an“ 2 ‘ § ‘ q ----— branch-92¢ "“--\-_ - ~ ----trunk -N‘ ‘ ........ uurk-SOII 40 L 1 1 _[ _L I | 0.05 0.075 0.1 0.125 0.15 0.17 stand density (#/m2) Figure 5-6 Modeled backscattering of different 5 forest components: leaf size (a); LAI (b); Branch radius (c); Branch density (d); DB (e); Stand density (1); Tree height (g). H modeled sigmo (dB) modeled sing (dB) -2 -——-———~ _2 1‘ -___. - - _ -_ _- - WW. ,I I ’4 d (a) A '4 _I (b) I m I g I -5 — c -6 d I E I -8 - o o g 8 I O O O . ' _ . . O I o .6” . o o z .8 o 9. :. . I 4049.‘ O 3-10“ .’09 . I ‘.90 E 0.3 0 -12 d o -12 - . I -14 I I I I -14 r I I _iI 0 2 4 6 8 12 o 5 1o 15 25 30 LAI tree height (m) -2 ‘ -2 _ ...... (C) I (d) ‘4 ‘ I -4 ' I 3 i 'o -6 I . g 5 - I E 8 I '5 -8 - - , . . ’ : :°.:°:.-° ’2 r 2’ .2: -102 ' ° 9 o, , -§-10- °° .0. k . E ’9’ . -12 “ O z -12 —I . £ '14 l l I I -14 1 I 1 1 I 0 5 10 15 20 0 10 20 30 4O 50 stem height (m) DBH (cm) -2 ..... _ (e) A -4 I m 3 o ‘6 3 E 2" to -8 r o . E o .3 .0... o o o g '10 “ O Q 0 . CE) ° 0... ’ -12 _ O -14 I I O 500 1000 1500 stand density (#lha) Figure 5-7 Backscattering simulation based on ground-measured forest parameters: LAI (a), tree height (b), trunk height (c), DBH (d), and stand density (e). 139 Figure 5-8 (a) A '8 d ’ o m 0 3 ’ o ’ : o '9 d ’ ’ 0 o E ’ o o m '5 -10 — . ’ -o o ’ o 2 O .8 -11— . o o O o E -12 _ O '13 l l I -14 -12 -1O -8 JERS-1 sing (dB) 3 _w_, m W O 2 _ 0)) . O _ 1 . . t . g Q .. I E 0‘ """"""""""" . “‘“3'; “““ “.3“? """"" I 2 ° 0 o o o I -1 O . O . -2 _ I -3 l l fl I -14 -12 -10 -8 modeled sing (dB) Comparison of J ERS-l SAR observed and modeled backscattering coefficients: scatterplot of backscattering coefficients (a) and residuals (b). The 1:1 line is also drawn in (a). 140 40 x stem_H R2 = 0_72 El DBH R2 = 0.84 35- 30- 25‘ (m) and DBH (cm) stem_H 8 0 5 1 0 1 5 20 25 30 tree_H (m) Figure 5-9 Scatterplot of stem height and DBH to tree height measured at study sites. Figure 5-10 Modeled tree height distribution in the watershed. 141 Density 600 600+ Figure 5-11 Modeled stand density distribution in the watershed. Figure 5-12 Error distribution of model inversion. 142 modeled stand density (#lha) Figure 5-13 -¥ N N 01 O 01 _x O l modeled tree height (m) 01 O 2000 1500 1000 500 Scatterplots of modeled and measured tree height (a) and stand density (b) l .0 O.“ 90. RMSE=4.6 meter (with outlier) RMSE=3.8 meter (without outlier) I I I I I 5 10 15 20 25 measured tree height (m) 30 (a) l RMSE=320 #/ha (with outlier) RMSE=319 #/ha (without outlier) 0. O ‘:.. ”O O. I I I 500 1000 1500 2000 measured stand density (#lha) (b) at all study sites. 143 25 ~—— — ~ —— — 7~ ~— — E, 20 -— — — — —— — ~— — i” 15 _ h _ ___. _ _ ‘3; ._ lmeasured g — _" T T ‘ “ . Dmodeled 3 . , , i: l 7 I I I dry mixed pine dry moist dipt trans ever ever (a) 1000 7'7 800 S I'—.—.. s a _ I g 600 .. _- .. __ __ lmeasured 2 f I. __ Elmodeled % 400 ~ __ is- _ U . , '5 200 " —j. "’7 O I If I I I4 I dry mixed pine dry moist dipt trans ever ever (b) Figure 5-14 Comparison of modeled and measured tree height (a) and stand density (b) in different forest types. 144 Chapter 6 Aboveground Woody Biomass Estimation with Microwave and Optical Remotely Sensed Data 6.1 Introduction Most of the world’s terrestrial biomass (SO-92%) is contained in forest ecosystems (Beaudoin et al. 1994). Aboveground forest biomass, or so-called biomass density, is of great importance in understanding the state and dynamics of ecosystems and their interaction with global carbon cycles/climate change studies because of its ability to lock up carbon and its potential for carbon exchange with the atmosphere through forest destruction and regrowth (Luckman et al. 1998). In this study, aboveground biomass represents the total woody biomass of trees of diameter 10cm or larger, including twigs, branches, trunks, and bark (Brown and Lugo, 1992). The amount of green leaves could be represented by LAI, which has been deeply studied with optical remote sensing techniques (Peterson et a1. 1987; Baret and Guyot 1991; Qi et al. 1995, 2000; Turner et al. 1999). In the past decades, forest biomass mapping from SAR data has become one of the most promising applications of radar remote sensing to vegetation studies. A number of studies have been done to estimate aboveground forest biomass using SAR remotely sensed data. One of the most common approaches is to empirically relate SAR backscattering intensity to biomass with ground measurements. Le Toan et al. (1992) found that there 145 was a good correlation between SAR intensity and forest biomass for lower frequency such as P- and L-band data. C-band data primarily interacted with the crown layer (Saatchi and McDonald 1997) and rapid saturation of the C-band backscattering coefficient occurred with the increase in forest biomass (Dobson et al. 1992). Cross polarized backscatter (HV) was found to have the highest correlation with forest biomass. HH is related to both trunk and crown biomass, whereas W and particularly HV returns are linked to crown biomass (Beaudoin et al. 1994). Luckman et a1. (1997) found that lower frequency (L-band) SAR imagery could be used to discriminate different levels of forest biomass up to 110 ton/ha. With the effect of topography, the biomass threshold, above which there appears to be no further increase in L—band radar backscatter, was around 60 tons per hectare (t/ha), and the error is estimated to be around 30% of the biomass value (Luckman 1998). Other studies (Le Toan et al. 1992, Dobson et al. 1992) using P- and L—band data in temperate forests indicated that it was possible to estimate biomass up to 200 ton/ha. With sensors at very long wavelengths (>1m) such as VHF SAR data, the backscattering coefficient is strongly correlated to characteristics of the tree trunk layer and the signal saturation is not observed up to 360 tons/ha (Melon et al. 2001). However, the empirical models did not add any physical explanation of the processes that drive the changes in backscattering. The biomass retrieval by means of relating biomass to SAR backscattering is shown to be ill-posed (Dobson et al. 1995). These relationships are highly affected by the structural properties of the forest such as species, age, height, 146 diameter, and tree component orientation distributions (Dobson et al. 1995), and other factors as topography, soil moisture, history, and local disturbances. It is also problematic when expanding the measured commercial biomass to the total aboveground biomass. Several canopy scattering models were successfully developed to simulate radar backscattering in temperate forests when these structural parameters were taken into account (Karam et a1. 1995 ; Ulaby et al. 1990; Sun et al. 1991). These models are a theoretical approximation of forest scattering based on the first or second resolution of radiative transfer functions (RTF). Among these models, the soil surface was simply assumed as a continuous and slightly rough dielectric surface. In forests with higher biomass, the soil backscattering is low and this simplification is acceptable. However, in sparse forests such as highly degraded or young secondary forests, the contribution of soil backscattering is under-estimated which introduces high error in biomass estimation. Due to the complicated natures of the input parameters, it is not a straightforward task to develop an effective inversion algorithm of these physical models for the purpose of biomass retrieval. Pierce et al. (1994) applied an artificial neural network to fulfill the model inversion of forest biophysical attributes. Ranson et al. (1997) used a combination of forest succession model and a canopy scattering model to develop the relationships between biomass and backscattering coefficients. The approach produced similar results with ground measurements and observed SAR data when aboveground biomass is less than 150 ton/ha, in the same level of direct empirical biomass retrieval. 147 When applied to SAR data, both the empirical and microwave RTF model—based biomass retrieval approaches only work on forest with sparse to medium biomass such as temperate coniferous or broadleaf forests. In dense tropical forests, because of the strong attenuation from the leaf layer, the radar backscattering is no longer sensitive to biomass when the forest reaches the threshold biomass level. If the contribution of leaves to the total backscattering can be quantified, the accuracy of woody biomass estimation could be improved at a much higher level. Optical remotely sensed data is mostly sensitive to the green leaf properties of the forest canopy. In optical remote sensing, the reflected energy of the Earth surface is determined by surface physical properties such as strong chlorophyll absorption and low reflectance (and, transmittance) in visible bands, and high reflectance due to internal leaf structures in near infrared (NIR). Optical remote sensing has been widely used to derive information of vegetation properties such as fractional cover and green leaf area index (J asinski and Eagleson 1990; Qi et al. 2000). A synergistic use of optical and microwave remote sensing could quantitatively evaluate the effect of leaf canopy to SAR backscattering and, therefore, could enhance the availability of biomass estimation. In this chapter, several methods were used to estimate aboveground biomass in the study area. Firstly, a simple regression model was built from the relationship between JERS-l SAR backscattering coefficients and ground-measured biomass. Then, with the forest structural parameters estimated in Chapter 5, the allometric equations were also applied to calculate biomass distribution in the study area. Finally, the backscattering 148 contribution of leaves and their attenuation effects to other forest components were calculated based on both JERS-l VNIR data and the microwave/optical synergistic model built in Chapter 3. The modeled backscattering with leaf compensation was then compared with ground measurements to estimation biomass distribution in the study area. 6.2 Biomass Estimation with a Simple Regression Method In the past studies, a simple 0'0 ~biomass regression model was often built to estimate biomass with SAR data (Luckman et al. 1997, Hashim and Kadir 1999, Santos, et al. 2003). In this section, the 0'0 in JERS-l SAR image at each study site was retrieved and the 0'0 ~biomass at all study sites was plotted (Figure 6-1). It was expressed as: 035R“ 2 ——13.838 + 0.963 * ln(bi0mass — 5.487) (6-1) Here the biomass is in the unit of ton/ha. The 0'0 was logarithmically related to biomass. At degree of freedom = 29 (32 study sites — 3 model coefficients), the x2 of curve fitting was 29.15, much smaller than the critical [3.0529 (46.19) at confidence level of 95%, indicating that the curve fitting in Figure 6-1 was valid and the 0'0 and biomass were logarithmically related. With Eq.6-1, the biomass distribution in the study area can be estimated with J ERS-l data (Figure 6-2). Unlike the forest fractional cover and leaf area index distribution, the woody biomass distribution estimated in this regression model did not agree with the forest type distribution. In the east of the watershed is Chiang Mai city and the forests are inevitably disturbed by human activities at a higher intensity. The Mae Chaem Town is 149 the only major human settlements in the middle and east of the watershed and the forests are less disturbed. This could be the possible reason that there had higher biomass distribution in the west than that in the east. However, in lack of a second set of ground measurements, it was impossible to validate this map. Moreover, the west part is in the near-range of J ERS-l SAR sensor whereas the east part is in far—range. Although the J ERS-l SAR data applied in this study has been calibrated and the pixels have been changed from slant resolution to ground resolution, the mis-calibration in this high mountainous area could also be a possible reason for the east-west variation of biomass distribution. 6.3 Biomass Estimation with Compensation of leaf Attenuation With the leaf area index (LAI) retrieved from JERS-l VNIR data, the attenuation effect of leaf layer on other forest components could be quantified in the microwave/optical synergistic canopy scattering model. Figure 6-3 is the LAI image created using the method described in Chapter 5. The non-forest open areas are masked out in the analysis. Since the J ERS-l data was acquired in the dry season, the dry dipterocarps and mixed deciduous forests had low LAI values (around 1.0 or lower) while the evergreen forests had high LAI values (4.0 or higher). A microwave/optical canopy scattering model was developed in Chapter 5 to simulate the backscattering from tropical forests. As shown in Figure 6-4a, there is a logarithm tic relationship between biomass and total modeled backscattering coefficients at the study sites. The modeled total backscattering scattering increases rapidly with biomass, then 150 quickly reaches the saturation point at biomass around 100 ton/ha, consistent with the past studies (Le Toan et al. 1992, Dobson et al. 1992, Luckman et al. 1997). The backscattering from each component of forests, however, changes differently with the increase of biomass (Figure 6-4b). In L-band HH polarization, the backscattering from branches plays the most important role and its variation with biomass is similar with the total backscattering. Trunk-ground double bounce is the second important backscattering component when biomass is lower. It decreases rapidly with higher biomass and reaches infinity (-80dB in the model) when biomass > 400 ton/ha. Leaf volume scattering also increases with biomass, but the residual is very high because leaf amount is not directly related to woody biomass, especially in the dry season in the study area. Branch-ground interaction, trunk volume scattering, and leaf volume scattering also decreases with higher biomass but with a lower rate than that of the trunk-ground interaction. With the microwave/optical canopy scattering model and the tree height map deve10ped in Chapter 5, the attenuation factor, I, at each study site could be calculated from Eq.5-13 to 5-15 when LAI was known. The attenuation from leaf layer can also be compensated and the scattering from the woody components (branches and trunks) be modeled. The leaf-compensated backscattering also had a logarithmic relationship with biomass. Figure 6-5 showed the logarithmic regression between biomass and J ERS-l SAR measured, modeled, and modeled after attenuation compensation in L-band and HH polarization (same system parameters as J ERS-l SAR). The JERS-l backscattering increases quickly 151 with biomass, then it slows down at biomass of around 100 ton/ha and tends to saturate. The modeled backscattering is about ldB lower than JERS-l, but the trend with biomass is similar. The modeled backscattering with leaf attenuation compensation is more sensitive to biomass. With limited study sites, the increase of backscattering does not slow down until biomass reaches 200 ton/ha. There is no obvious saturation in the modeled backscattering (no-leaf) in Figure 6—5. At lower biomass, the green leaves have significant contribution to the total backscattering and, therefore, the modeled backscattering is higher than the one after leaf attenuation compensation. When biomass is higher than 300 ton/ha, the leaf attenuation to the woody components (branches, trunks) becomes significant and therefore, the modeled backscattering is lower than the one after leaf attenuation compensation. The coefficients of these curve fittings and their statistical tests are listed in Table 6-1. For a chi-square test, with degree of freedom=29 (32 - 3 coefficients in the model), the critical value, 1505,29 , is 46.19 at 95% confidence level. The x2 values of 03A,, and 030d 6, in Table 6-1 are much smaller than 150529 and, therefore, the curve fitting lines are valid in Figure 6-5. The x2 values of agwde,("o_,wf) is very close to the 130529. One possible reason is that when the leaf attenuation is compensated, the topographic effect on the SAR backscattering and its interaction with ground surface becomes more significant, which introduces higher uncertainty to 0310de,(,,0_,mf) . 152 Figure 6-6 is the leaf scattering (Figure 6-4a) and leaf attenuation (Figure 6-4b) images in the study area in the unit of dB. The leaf scattering component (0780]) is the combined intensity of leaf volume scattering and leaf-ground interaction. The leaf attenuation factor (1:) is the attenuation to the radar signals reaching or scattering from the branches, trunks, and soil ground. The leaf scattering ranged from —32 dB in dry dipterocarps to —10 dB in tropical evergreen forests. The leaf attenuation (r) ranged from 0.99 in dry dipterocarps to 0.19 in tropical evergreen forests. The leaf scattering and its attenuation effects are positively correlated to LA] values. When LAI increased in the study areas, both the leaf scattering and its attenuation to woody structures increased. As a result, the value of attenuation factor (I) was lower. Dry dipterocarp forests have lowest leaf scattering and highest attenuation factor (r) values. In mixed deciduous forests, the leaf scattering is higher and t is lower. Tropical evergreen forests have highest leaf scattering and lowest 1: values. With JERS-l SAR data, LAI data, and the microwave/optical synergistic canopy scattering model, the leaf scattering and its attenuation to woody forests could be calculated. The scattering coefficient (a ) from woody forests (branches, trunks) is woody then quantified by subtracting leaf scattering (am) from the total backscattering (CSAR) then compensated for the two-way leaf attenuation (t). It is the combined contribution of volume scattering from branches and trunks and their interaction with ground surface. In power unit, 0' could be expressed as: woody _ USA}? —Uleaf (6 5) woody _ z_2 - 0' 153 The woody scattering distribution in the study area was shown in Figure 6-7. It is obvious that after leaf attenuation compensation, the radar scattering from woody forests (branches and trunks) is more significantly influenced by topography in the study area. Tropical evergreen forests have higher scattering from woody structures, but the values vary greatly because of the topographic effect. With the regression model described in Figure 6-5 and Table 6—1, the biomass distribution was estimated with the relationship between the woody forest scattering and biomass (Figure 6-8). The topographic effect in tropical evergreen forests was overwhelming so that the estimated biomass in these forests was even lower than dry dipterocarps and mixed deciduous forests. From both Figure 6-2 and Figure 6-8, the regression models could not successfully estimate the biomass in the study area, especially in tropical evergreen forests where the topographic effects were significant. The results are either overestimated or underestimated in different forests. There was always a trend of increasing biomass from west to east of the study area. 6.4 Biomass Estimation with Synergistic Model and Allometric Equations The modeled tree height and stand density distributions in the study area were described in Chapter 5. A least-square optimization technique was applied in model inversion while the leaf scattering and its attenuation to woody forests (branches, trunks) were quantified with J ERS-l VNIR data. Also, as described in both Chapter 2 and Chapter 5, the diameter at breast height (DBH) was linearly correlated with tree height. With the tree 154 height distribution map in Chapter 5, the DBH distribution was modeled (Figure 6-9). Similar as the tree height distribution, most of dry dipterocarps have lower DBH values (around 15-20 centimeters). Mixed deciduous forests have a wider range of DBH values between 15 -40 centimeters. The tropical evergreen forests have the highest heterogeneity. Some of the areas in tropical evergreen forests could have DBH values larger than 45 centimeters while in some isolated areas they are smaller than 10 centimeters. There is also a trend of DBH values increasing from west to east of the watershed. The trend is not obvious in other forest types. From the modeled tree height, stand density, and DBH distributions, the aboveground forest woody biomass in the watershed could be calculated with the allometric equations described in Chapter 2. In Figure 6-10, the dry dipterocarp forests have the lowest biomass (SO-150 ton/ha). There are also some isolated areas with biomass less than 50 ton/ha. The biomass in the areas close to human settlements is lower than the ones far away. Mixed deciduous forests have higher biomass ranging from 50 to 200 ton/ha. In some isolated areas it could reach 400 ton/ha. The biomass distribution in tropical evergreen forests has the highest heterogeneity than any other forest types. The biomass values range from 50 ton/ha to more than 400 ton/ha. In some areas the biomass estimation is saturated. In some isolated areas, the biomass distribution is even less than 50 ton/ha. The biomass estimation in tropical evergreen forests is questionable because of two reasons: first, tropical evergreen forests locate atop of mountains in the study area. In some areas, the topographic relief is so high and the 155 slopes are so steep that the topographic effect to the J ERS-l SAR image cannot be corrected. It introduces high error of model inversion in the areas. Second, the modeled stand density in tropical evergreen forests saturates at’around 500 trees/ha (Chapter 5), and therefore, the accuracy of the biomass estimation in these areas are lower than dry dipterocarps and mixed deciduous forests. The modeled biomass values at the study sites were compared with ground-measured values (Figure 6-11). The modeled and measured biomass values scattered along the 1:1 line. The biomass values at most of the study sites were less than 200 ton/ha. Both the modeled and measured biomass values at moist evergreen forests were higher than 350 ton/ha. There were one pine transition site and one dry evergreen site at which the modeled biomass was underestimated. The site (SZ-14) was an outlier at which the modeled biomass was highly overestimated. This overestimation came from the model saturation. The site was also an outlier in the scatterplot of modeled and measured tree height at study sites in Chapter 5 (Figure 5-1 3a). There was no study site with biomass between 200 to 350 ton/ha in both modeled and measured methods. The total root-mean- square error (RMSE) was 121 ton/ha with this outlier (82-14), and. 88 ton/ha when the outlier was removed. When compared in different forest types (Figure 6—12), the modeled biomass was obviously overestimated at dry dipterocarps study sites because of the overestimation in the modeled tree height at study sites. In mixed deciduous, pine transition, and dry evergreen forests (when the outlier S2-l4 was removed), the average values of modeled 156 biomass fitted well with measured values. However, the standard error of modeled biomass in pine transition was very high, indicating high heterogeneity in the zone. Even when the outlier (S2-14) was removed, the standard error of dry evergreen was still slightly higher than that of dry dipteorcarps and mixed deciduous. Both the modeled and measured biomass values at moist evergreen forest sites were higher than 400 ton/ha. The model tended to saturate and therefore the standard error was very low. 6.5 Conclusions and Discussions In this chapter, several methods for biomass estimation with microwave and optical remotely sensed data were examined. Firstly, a simple regression model was developed with the observed J ERS-l SAR backscattering coefficients and ground-measured biomass in the study sites. The model indicated low biomass in open areas and dry dipterocarp forests close to the human settlements, and high biomass values in mixed deciduous forests. However, due to the severe topographic effect to the JERS-l SAR data, the estimated biomass in tropical evergreen forests was low and varied greatly. There was an obvious trend of increasing biomass from east to west of the study area. The possible reason comes from the mis-calibration of JERS-l SAR sensor. With the limited study sites and biomass measurements, it was impossible to validate the biomass map derived with this method. Secondly, with the J ERS-l VNIR optical data and the microwave/optical synergistic canopy scattering model developed in Chapter 5, the volume scattering from leaf layer in forests and its interaction with ground surface was quantified, and its attenuation factor 157 (1:) was calculated. Then the scattering from woody forests (branches, trunks) was modeled by subtracting the modeled leaf scattering from the observed JERS-l SAR backscattering coefficients and compensating the attenuation from leaf layers. According to the relationship between the woody scattering and biomass over the study sites, the woody scattering was more sensitive to biomass when the attenuation from leaf layer was compensated. A similar regression model was built based on the relationship between the modeled woody scattering and ground-measured biomass over the study sites. It was applied to estimate the biomass distribution in the study area. However, when the leaf scattering was removed and its attenuation to the woody forests compensated, the backscattering coefficients were more severely affected by topographic variation. The modeled biomass also revealed a trend of increasing biomass from east to west of the study area. Similarly, the biomass map in this method cannot be validated. Thirdly, with the tree height and stand density distributions with the microwave/optical synergistic canopy scattering model in Chapter 5, the DBH distribution was estimated and the biomass was calculated with the allometric equations. The biomass map in this method clearly showed the lower biomass distribution in dry dipterocarp forests (SO-150 ton/ha, less than 100 ton/ha in most areas), medium distribution in mixed deciduous forests (SO-200 ton/ha, greater than 100 ton/ha in most areas), and high biomass distribution in tropical evergreen forests (50-500 ton/ha and higher). Tropical evergreen forests showed a high heterogeneity in biomass distribution. It possessed the lowest biomass in the areas with high relief and steep slopes where the topographic effects cannot be corrected. As mentioned in Chapter 5, the stand density estimation with model 158 inversion did not work in tropical evergreen forests and saturated at around 500 trees/ha. Therefore, the biomass estimation in tropical evergreen forests tends to saturate. The estimated biomass in the third method was compared with the ground—measured biomass in all study sites. The root-mean-square error of the biomass estimation in the study sites was 88 ton/ha. The biomass was overestimated in dry dipterocarp forests. The model worked well in mixed deciduous forest. But the pine transition and dry evergreen forests, which belong to tropical evergreen forests in the forest type map, had high standard error, indicating high heterogeneity in these forests. In conclusion, the synergistic use of microwave and optical remotely sensed data in a canopy scattering model provided a new approach of biomass estimation in tropical forests. However, this method is highly limited by the poor quality of SAR data and geometric mismatch between SAR and DEM data. In mountainous areas where the topographic effect to the SAR data cannot be corrected, the biomass estimation is very questionable. The model could be applied to estimate forest biomass up to 200 ton/ha. In forests with higher densities (biomass higher than 400 ton/ha), the SAR signal begins to be saturated and is no longer sensitive to the biomass increments. The accuracy of ground measurements is also critical to the validation of this method. With limited study sites, the outlier of one site could result in 33 ton/ha of RMSE of biomass estimation in the area. Moreover, no study sites with biomass between 200 to 350 ton/ha were visited. As a result, it is impossible to validate the biomass estimation in 159 this range. Since the ground-measured biomass at study sites with point-quadrant method tend to be underestimated in heterogeneous forests, a better method for ground measurements is needed in the future research. 160 6.6 References Baret F. and Guyot, G. (1991). Potentials and limits of vegetation indices for Lai and APAR assessment. Remote Sensing of Environment: 35, 161-173. Beaudoin, A., Le Toan, T., Goze, S., Nezry, E., Lopes, A. and Mougin, E. (1994). Retrieval of forest biomass from SAR data. International Journal of Remote Sensing: 15(14), 2777-2796. Brown, S. and Lugo, A. E. (1992). Aboveground biomass estimates for tropical moist forests of the Brazilian Amazon. INTERCIENCLA: 17(1), 8-18. Brown, S., Gillespie, A. J. R. and Lugo, A. E. (1989). Biomass estimation methods for tropical forests with applications to forest inventory data. Forest Science: 35(4), 881-902. Dobson, M. C., Ulaby, F. T., Le Toan, T., Beaudoin, A. and Kasischke, E. S. (1992). Dependence of radar backscatter on coniferous forest biomass. IEEE Transactions on Geoscience and Remote Sensing: 30, 412-415. Dobson, M. C., Ulaby, F. T., Pierce, L. E., Sharick, T. L., Bergen, K. M., kellndorfer, J ., Kendra, J. R., Li, B, Lin, Y. C., Nashashibi, A., Sarabandi, K. and Siqueira, P. (1995). Estimation of forest biophysical characteristics in Northern Michigan with SIR-C/X—SAR. IEEE Transactions on Geoscience and Remote Sensing: 33, 877- 895. Frazer, G. W., Canham, C. D. and Lertzman, K. P. (1999). Gap Light Analyzer (GLA), Version 2.0: Imaging software to extract canopy structure and gap light transmission indices from true-color fisheye photographs, users manual and program documentation. Copyright © 1999: Simon Fraser University, Burnaby, British Columbia, Canada, and the Institute of Ecosystem Studies, Millbrook, New York, USA. Hashim M. and Kadir, W. H. W. (1999). Comparison of JERS-l and Radarsat synthetic radar data for mapping mangrove and its biomass. In the 20th Asian Conference on Remote Sensing, Nov. 22-25, 1999, Hong Kong, China. Poster session 1, 1-5. Henderson, F. M. and Lewis, A. J. (1998). Principles and applications of imaging RADAR, Manual of remote sensing, vol.2, third edition. John Wiley &Sons, Inc. Le Toan, T., Beaudoin, A. and Guyon, D. (1992). Relating forest biomass to SAR data. IEEE Transactions on Geoscience and Remote Sensing: 30, 403-411. Luckman A. J. (1997). The effects of topography on mechanisms of radar backscatter from a coniferous forest and upland pasture. IEEE Trgnjsactions on Geoscience and Remote Sensing: 36(5), 1830-1834. 161 Luckman, A., Baker, J ., Honzak, M. and Lucas, R. (1998). Tropical forest biomass density estimation using JERS-l SAR: seasonal variation, confidence limits, and application to image mosaics. Remote Sensing of Environment: 63, 126-139. Luckman, A., Baker, J ., Kuplich, T. M., Yanasse, C. C. F. and Frery, A. C. (1997). A Study of the relationship between radar backscatter and regenerating tropical forest biomass for spacebome SAR instruments. Remote Sensing of Environment: 60, 1-13. Melon, P., Martinez, J. M., Le Toan, T., Ulander, L. M. H. and Beaudoin, A. (2001). On the retrieving of forest stem volume from VHF SAR data: observation and modeling. IEEE Transactions on Geoscience and Remote Sensing: 39(11), 2364- 2372. Peterson, D. L., Spanner, M. A., Running, S. W. and Teuber, K. B. (1987). Relationship of thematic mapper simulator data to leaf area index of temperate coniferous forest. Remote Sensing of Environment: 22, 323-341. Pierce, L. E., Sarabandi, K. S. and Ulaby, F. T. (1994). Application of an artificial neural network in canopy scattering inversion. International Journal of Remote Sensing: 15(16), 3263-3270. Qi, J ., Cabot, F., Moran, M. S. and Dedieu, G. (1995). Biophysical parameter estimations using multidirectional spectral measurements. Remote Sensing of Environment: 54, 71-83. Qi, J ., Chehbouni, A., Huete, A. R. and Kerr, Y. (1994). A modified soil adjusted vegetation index (MSAVI). Remote Sensing of Environment: 48, 119-126. Qi, J ., Kerr, Y. H., Moran, M. S., Weltz, M., Huete, A. R., Sorooshian, S. and Bryant, R. (2000). Leaf area index estimates using remotely sensed data and BRDF models in a semiarid region. Remote Sensing of Environment: 73: 18-30. Ranson, K. J. and Sun G. (1997). An evaluation of AIRSAR and SIR-C/X-SAR images for mapping northern forest attributes in Maine, USA. Remote Sensing of Environment: 59, 203—222. Saatchi, S. S. and McDonald, K. C. (1997). Coherent effects on microwave backscattering models for forest canopies. IEEE Transactions on Geoscience and Remote Sensing: 35, 1032-1044. Santos, J. R., Freitas, C. C., Araujo, J. S., Dutra, L. V., Mura, J. C., Gama, F. F., Soler, L. S., and Sant’Anna S. J. S. (2003). Airborne P-band SAR applied to the aboveground biomass studies in Brazilian Tropical Rainforest. Remote Sensing of Environment: 87, 482-493. 162 Sun, G., Simonett, D. and Strahler, A. (1991). A radar backscatter model for discontinuous coniferous forests. IEEE Transactions on Geoscience and Remote Sensing: 29, 639-650. Thailand Land Use and Land Cover Change Case Study (1997). Southeast Asia, IHDP, IGBP, WCRP program. Turner, D. R, Cohen, W. B., Kennedy, R. E., Fassnacht, K. S. and Briggs, J. M. (1999). Relationships between leaf area index and Landsat TM spectral vegetation indices across three temperate zone sites. Remote Sensing of Environment: 70, 52-68. Ulaby, F .T., Sarabandi, K., Mcdonald, K., Whitt, M. and Dobson, M. C. (1990). Michigan microwave canopy scattering model. International Journal of Remote Sensing: 11(2), 1223-1253. 163 Table 6-1 Coefficients of 0'0 ~biomass logarithmic curve fitting and their statistic tests. 0' 321R Grand e! Grind el(no-leaf) m0 -14.22 -14.43 -19.09 ml 1.04 0.93 1.73 m2 -3.29 -11.10 -7.14 X2 24.22 27.58 46.08 Correlation coefficient (R) 0.75 0.79 0.79 Equation 0'0 = m0 + ml * ln(bi0mass + m2) 164 JERS sigmHH 0 100 200 300 400 500 600 biomass Figure 6-1 Scatterplot and curve fitting of 035R“ ~biomass at all study sites. 300 >300 Figure 6—2 Biomass distribution estimated with a simple 0'3““ ~biomass regression model. 165 Figure 6-3 The LA] distribution derived from J ERS-l VNIR data in the study area. 166 -7 -8 T . .. . . I 90 o g '9 ‘ a? . . .9 - - ‘5 m 10 ’ . . '5 o 2 -11 - .9 a.) O s 12 _ E - O -13 - —14 I I 0 200 400 600 biomass (a) 0 _ x A -10 8° 35 my XQ 0° :0 0g >$< 3 _20 _ o E U) '71 -30 ~ 3 x oleaf % _40 _ + Aleaf-soil g x xbranch A _50 - + xbranch—soil .5 otrunk -60 ' 0 I I +trunk-soil 0 200 400 600 biomass (b) Figure 6-4 Relationships between ground-measured biomass and modeled backscattering coefficients of forests (a) and their components (b). 167 -1 3 . . - backscattering coefficient (dB) I . 444/ A-7JERS=18AR . '5' —-—- Model '15'I ------- ModeunoIean -16’ 160 260 360 460 560 600 biomass (ton/ha) Figure 6-5 Logarithmic curve fitting of SAR observed and modeled backscattering coefficients with biomass. Figure 6-6 Leaf scattering (a) and its attenuation to the woody forests (b). 168 Figure 6—7 Woody scattering in the study area. Figure 6-8 Biomass distribution with woody forest scattering in a regression model. 169 DBH (cm) <15 15 Figure 6-9 Modeled DBH distribution in the watershed. 300 >300 Figure 6-10 Forest biomass distribution from model inversion in the watershed. 170 600 RMSE=121 ton/ha (with outlier) 75‘ RMSE=88 ton/ha (without outlier) E 500 — 9 g I ’ o a 70’ 400 — 82-14 ’ ° to E 2 300 ~ .0 g 200 ~ .3 ‘ , . g o .9 o O 100 ~ ‘ o.’ e 14‘ O I I | I l 0 100 200 300 400 500 600 measured biomass (ton/ha) Figure 6-11 Scatterplot of modeled and measured biomass at study sites. 500 I I I I a modeled A 400 T I measured 3 . standard error 5 300 — 8 200 - E f " .9 If. _D :14 . . I, 100— I I dl')’ mixed pine dry moist dipt trans ever ever Figure 6-12 Average values of measured and modeled biomass (without outlier) and standard error of the modeled biomass in each forest type. 171 Chapter 7 Conclusions and Future Envisions 7.1 Concluding Remarks The objective of this study is to quantitatively estimate biophysical attributes using both optical and microwave (SAR) remote sensing techniques in tropical forests. Several physical, semi-empirical, and empirical models were built to address this objective. The study area is tropical environment in Mae Chaem Watershed, ChiangMai, Thailand. The field measurements were described in Chapter 2. The biophysical attributes estimated in this study include forest fractional cover (Chapter 3), leaf area index (Chapter 4), tree height and stand density (Chapter 5), and aboveground biomass (Chapter 6). A linear unmixing model was built to estimate forest fractional cover (Chapter 3). Instead of using the spectral reflectance in the linear unmixing models that were commonly applied in the past studies, a vegetation index — MSA VI, which is most linearly related to the green leaf abundance in tropical forests when LA] is less than 4.0 — was applied in this study. Only two components were assumed in each pixel of the remote sensing imagery: tree canopy and open area. The forest fractional cover thus represents the coverage of the forest canopy in each pixel. A forest fractional cover map was built with a Landsat ETM+ image acquired in the dry season in the study area. It was validated with both ground- measured fractional cover at 32 study sites (R2=0.76) and high-resolution (1 meter) IKONOS calculated fractional cover data at 400 randomly selected locations (R2=O.7O). The fractional cover map was also adjusted to the wet season in which all forest types 172 were flourishing and the seasonal variation of deciduous species in the forests was reduced. Leaf area index is highly correlated to forest fractional cover. A modified Gaussian regression model was built to estimate leaf area index with forest fractional cover (Chapter 4). Since it is impossible to measure leaf area index with LAI—2000 at the study sites in the watershed, additional ground data were acquired in northern Michigan where several forest areas were measured at different seasons to represent the forest conditions in the study area in Thailand. Both LAI-2000 and fisheye photos were taken to measure leaf area index and forest fractional cover in these forests. The regression model was examined with a x2 goodness-of-fit test and then applied to the study area in Thailand. The leaf area index in both dry and wet seasons were mapped in the study area. A narrow zone, possibly the pine transition zone between mixed deciduous and evergreen forests was shown in the leaf area map in the dry season. The leaf area index in the wet season was much higher in the study area. But some forests not far from the large open areas (villages, agricultural areas) at lower elevation have very low values even in the wet season, indicating intense human disturbances in these forests. To further retrieve forest structural information and aboveground biomass, a microwave/optical synergistic radiative transfer model was built in Chapter 5. The volume scattering from woody components (branches, trunks) and their interaction with ground surface were simulated while the leaf scattering was quantified with leaf area index fi'om JERS-l VNIR optical data. The root-mean-square error (RMSE) of the 173 model was 0.94 dB compared with the J ERS—l SAR backscattering coefficients. Forest structural parameters, such as tree height and stand density, were estimated by model inversion with least-square-error optimization techniques. The error of model inversion in most of the study areas was less than 1 dB. In areas with high relief and steep slopes, the topographic effects on J ERS-l SAR backscattering coefficients cannot be corrected and the error of model inversion could be higher than 4 dB, introducing high error to the forest structural estimation of tree height and stand density. The RMSE of tree height estimation was 3.8 meter and that of stand density estimation was 299 trees/ha. The tree height in young secondary forests, mostly dry dipterocarp forests, was overestimated while the estimation in other forests fitted well with ground measurements. The stand density estimation in tropical evergreen forests did not work. It saturated at around 500 trees/ha. In accordance with the ground measurements, the modeled tree height was negatively correlated with the modeled stand density. Several methods of biomass estimation with J ERS-l SAR and VNIR data were examined in Chapter 6. A simple regression model was firstly built to estimate biomass from JERS- 1 SAR backscattering coefficients. The model was examined with a )8 test. It revealed a trend of increasing biomass distribution from east to west of the study area that was possibly the mis-calibration in the SAR data when converting slant—range to ground pixels. The estimated biomass of tropical evergreen forests was very low and varied greatly. With J ERS-l VNIR data, the leaf scattering and its attenuation to the woody forests (branches and trunks) were quantified. Then the woody forest scattering after the compensation of leaf attenuation was modeled. A new regression model was built to 174 estimate biomass with woody scattering in the study area. The woody forest scattering was more sensitive to the biomass, but the topographic effect to the scattering also became higher when the leaf attenuation was removed. The modeled biomass in tropical evergreen forests was even lower with higher heterogeneity. None of the two methods above could be validated in lack of additional ground measurements. The third method of biomass estimation was allometric equations. With the microwave/optical synergistic model and J ERS-l SAR and VNIR data, the tree height and stand density were estimated in Chapter 5. The DBH was estimated from its linear relation to tree height. Therefore, the biomass was able to be calculated with allometric equations. The biomass in this method was overestimated in dry dipterocarp forests, but fitted well with ground measurements in mixed deciduous forests. The biomass estimation in tropical evergreen forests is highly heterogeneous because of the topographic effect and the mis-estimation of stand density in Chapter 5. The root-mean-square error of the biomass estimation in the study sites was 88 ton/ha. In a summery, with both optical and microwave remote sensing imagery and radiative transfer model, the forest fractional cover, leaf area index, tree height and stand density, and aboveground biomass were mapped in the study area. Aside from forest fractional cover, each other biophysical attribute is mainly (such as leaf area index) or partially (such as tree height and stand density, and aboveground biomass) from the results in earlier chapters. The error propagation between these attributes will be studied in the near future. 175 7.2 Challenges The estimation of biophysical attributes in tropical forests mountainous areas was very challenging. The ground measurements are time and labor consuming and are often limited because of the physical difficulty of access to the study sites. Moreover, the ground measurements are not actual ground “truth” in many situations. For example, the ground-measured forest fractional cover and leaf area index with fisheye photo and LAI- 2000 are actually projected cover and foliage area index. They are the combined contribution from both green leaves and branches, trunks and any other non-open components. On the other hand, the fractional cover and leaf area index estimation with remotely sensed data are based on the spectral responses that are mostly sensitive to green leaves. The validation of remote sensing estimated fractional cover and leaf area index with ground measurements is thus biased. Another challenge of ground measurements is to measure the tree structural parameters such as tree height and stand density. The fixed-radius plot method often provides the accurate measure of these parameters on the study area, but is too time- and labor- consuming and not realistic in forests. The point-quadrant plot method is a more efficient way of ground measurements, but often results in high bias to the ground data. In the late succession of forests, there are usually many young trees that surround the old ones and, therefore, there is high possibility of picking up young trees in each quadrant. The measured tree height and stand density are thus underestimated. Assuming these data as ground “truth”, the estimated biophysical parameters from remotely sensed data are often 176 overestimated. In the future studies, the ground measurements with both methods at same study sites could be compared and a possible compensation model could be developed. Topographic effect in SAR images was the largest obstacle in microwave remote sensing. The quality of topographic correction to the SAR images depends on both the system characteristics of SAR sensor, the DEM data, and the target characteristics. In mountainous areas with high relief and steep slopes, the topographic effects cannot be corrected at all. The mis-correction results in high errors to the estimation of forest biophysical attributes. There are thousands of computing iterations in the model inversion and the canopy scattering model is called at each iteration for each pixel of the SAR and optical images. Therefore, the computation time is also very challenging in the estimation of forest structural parameters. A possible solution is to do the model inversion and to estimate biophysical attributes only at core locations in the study area. The attributes at other pixels can be estimated with linear or polynomial interpolation. When SAR and optical remote sensing data are applied synergistically, the acquisition of SAR and optical data at the same time becomes critical. Although quite a few optical sensors are available by now, there are only ERS-2, Radarsat, and Envisat together with some Airborne sensors (for example, AIRSAR) still operative. SAR sensors in both ERS- 2 and Radarsat are in C-band and their application in tropical forests is thus very limited. Fortunately, with high temporal resolution of optical data in the past years, it is possible 177 to develop leaf area index maps on an annual base, which could provide important information for the microwave/optical synergistic model. In the near future, the only opportunity to acquire SAR/optical imagery at the same time is the Advanced Land Observing Satellite (ALOS) designed by NASDA in Japan. It will be launched in 2004 boarding one radar sensor and two optical sensors: Phrase Array type L-band Synthetic Aperture Radar (PALSAR, a follow-up of J ERS-l SAR), Advanced Visible and Near Infrared Radiometer type-2 (AVNIR-2, a follow-up of J ERS— 1 VNIR), and Panchromatic Remote-sensing Instrument for Stereo Mapping (PRISM). While PALSAR and AVNIR-2 acquire SAR and optical data almost simultaneously, the PRISM sensor collects high-resolution elevation data that could improve the topographic correction on both SAR and Optical data. If launched successfully, ALOS data could provide valuable information in studies of forests of mountainous areas. 7.3 Potential Applications for Other Studies The results in this study could provide some quantitative information for the studies of land use / land cover change and human-environment interactions. The variation of the biophysical attributes distribution in the study area is related to the extent of human disturbance and forest recovery, together with the difference in forest types, soil texture, and topography. For example, the biophysical parameters had a pattern of rapid increase from downhill to uphill in Mae Chaem Watershed. This phenomenon revealed the effects of human activities on land use and land cover dynamics. The forests at downhill suffer from intense human disturbance of burning for agriculture and cutting for firewood. 178 Therefore, the values of the biophysical attributes are very low. At higher elevation, due to the geographical difficulty, clear-cuts for small area agriculture fields and selective logging for valuable trees become the major human activities. Consequently, the forest biophysical attributes have higher values. On the top of the mountain, the forests are only accessible to highland tribes whose activities are very limited and therefore, the biophysical distribution of these forests varies little. When applied to a larger scale, this research could also serve as a guide for ecosystem assessment and further decision-making for environment protection. For a long time, the binary forest/non-forest mapping has been applied in the assessment of carbon cycle and global climate change studies. However, different forests, such as mature forests, selectively logged forests, highly degraded and fragmented forests, and secondly regrowed forests which have different fractional cover and biomass, play different roles in carbon sequestration. The methods developed in this study can provide quantitative estimation of fractional cover and biomass in the forests that are important input parameters in carbon and climate models. With this information, we can assess the net carbon flux in forest ecosystems at higher accuracy. For future studies, the methods developed in this study could be applied to estimate biophysical attributes in other vegetated ecosystems like temperate forests, agricultural cropping systems, and rangelands. In each ecosystem, only the probability distribution functions of the vegetation components and their geophysical and biophysical parameters 179 in the synergistic radiative transfer model need to be re-evaluated. The model could also be modified and applied to SAR data at different polarizations and radar frequencies in the estimate of biophysical attributes in different ecosystems. 180 .1. Ear...) 33.3.... v .343 .1: : Eli H. W}.-. l. i HHS 15,/53:2 e D. n .u .N .52.... (‘12.. .t .. . 2w” J. : z: .4 : a}?